You are on page 1of 8

Applied Clay Science 80–81 (2013) 305–312

Contents lists available at ScienceDirect

Applied Clay Science


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

Adsorption–desorption characteristics of nitrate, phosphate and sulfate


on Mg–Al layered double hydroxide
A. Halajnia a,⁎, S. Oustan a, N. Najafi a, A.R. Khataee b, A. Lakzian c
a
Department of Soil Science, Faculty of Agriculture, University of Tabriz, 51664-16471, Tabriz, Iran
b
Research Laboratory of Advanced Water and Wastewater Treatment Processes, Department of Applied Chemistry, Faculty of Chemistry, University of Tabriz, Tabriz, Iran
c
Department of Soil Science, Faculty of Agriculture, Ferdowsi University of Mashhad, Iran

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

Article history: The adsorption–desorption behavior of nitrate on Mg–Al layered double hydroxide (4:1) was compared with
Received 21 December 2012 that of phosphate and sulfate as two common anions in soil solution. Based on the results, the kinetics of
Received in revised form 22 April 2013 anion adsorption on the LDH followed a pseudo-second order model. The adsorption process was found to
Accepted 6 May 2013
be exothermic for nitrate and endothermic for phosphate and sulfate. The adsorption data were best
Available online 25 May 2013
described by the Freundlich model for nitrate and by the Langmuir model for phosphate and sulfate. In this
Keywords:
study, the synthesized LDH exhibited higher adsorption rate and adsorption capacity for nitrate compared to
Adsorption capacity sulfate and phosphate. The maximum adsorption capacity values were 1.90, 0.28 and 0.13 mmol g −1 for
Adsorption selectivity nitrate, phosphate and sulfate, respectively. However, the Langmuir equation constants related to the adsorp-
Adsorption kinetics tion energy were found to be 0.210, 10.731 and 3.021 L mmol−1, respectively. The distribution coefficient
Hysteresis index (Kd) was higher for nitrate than phosphate and sulfate at high initial concentrations. Moreover, the values
FTIR spectroscopy of hysteresis index based on the Freundlich exponent were 97.5, 22.5 and 79.0% for nitrate, phosphate and
sulfate, respectively. The FTIR spectroscopy indicated a strong band related to the nitrate intercalated into
the interlayer space of the LDH.
© 2013 Elsevier B.V. All rights reserved.

1. Introduction belong to a group of lamellar non silicate compounds with positively


charge. LDHs are commonly represented by the formula of [M 2+1 − x
Due to the high mobility of nitrate and low anion exchange capac- M 3+x(OH)2] x+[A m−x/m].nH2O, where, M 2+ and M 3+ are divalent
ity of soils, this anion easily leaches through the soil profile. Nitrate and trivalent cations. The value of x is equal to M 3+/(M 2++ M 3+)
leaching is a serious concern for two reasons, decreasing nitrogen- and A m− is the charge balancing interlayer anion. In the majority of
fertilizer use efficiency (NUE) and water resources quality problems. cases the x values change between 0.10 and 0.33 (Forano et al.,
A significant amount of nitrogen fertilizer can be lost in the form of 2006). LDHs with positively charged brucite-like sheets and relatively
nitrate, so that the average nitrogen use efficiency is estimated less weak interlayer bonding exhibit excellent ability to capture different
than 50% (IAEA, 2008). Nitrate contamination of drinking water families of anions such as halides, non-metal oxyanions, anionic
from agricultural activities is a serious problem in many parts of the metal complexes, organic anions and anionic polymers (Forano et
world (Prakasa Rao and Puttanna, 2006). Increasing concerns to the al., 2006; Goh et al., 2008). It is generally believed that LDHs have
environmental impacts associated with nitrate leaching has led to greater affinities for anions with higher charge density. For common
the development of various methods to increase nitrogen fertilizer inorganic anions, the LDH selectivity decreases in the following
efficiency such as the use of slow release fertilizers, urease enzyme order: CO32− > HPO42− > SO42− > Cl − > NO3− (Das et al., 2006; Goh
and nitrification inhibitors and multi-stage application of these fertil- et al., 2008; Lv et al., 2008; Tezuka et al., 2004; You et al., 2001).
izers. Recently, use of layered double hydroxides (LDHs) as a slow Therefore, nitrate adsorption efficiency of LDHs is strongly reduced
release fertilizer or anionic exchanger has been proposed for increas- in the presence of other anions. For successful application of LDHs
ing NUE as well as nitrate buffering capacity of soils (Komarneni et al., to nitrate removal, these compounds should have high capacity and
2003; Torres-Dorante et al., 2008, 2009). Layered double hydroxides or selectivity for nitrate in the presence of other anions in complex
solutions such as soil solution. There are few studies about effective
LDHs with a high capacity and selectivity to nitrate adsorption.
Génin et al. (2001) illustrated the potential of Fe(II)–Fe(III) green
⁎ Corresponding author. Tel.: +98 511 878 9213.
E-mail addresses: halajnia@yahoo.com (A. Halajnia), oustan@hotmail.com
rust for reducing nitrate in soil solution. Tezuka et al. (2004, 2005)
(S. Oustan), n-najafi@tabrizu.ac.ir (N. Najafi), a_khataee@tabrizu.ac.ir (A.R. Khataee), found that Ni–Fe-LDH (4:1) had a high selectivity for nitrate com-
alakzian@yahoo.com (A. Lakzian). pared with Mg–Al, Co–Fe and Mg–Fe because of appropriate basal

0169-1317/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.clay.2013.05.002
306 A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312

spacing of the brucite-like layer. They reported that despite the great- 3. Results and discussion
er theoretical exchange capacity, Mg–Al-LDH (3:1) adsorbed a very
small amount of nitrate from seawater. However, Islam and Patel 3.1. Physical and chemical properties
(2009) showed that Mg–Al-LDH (3:1) efficiently adsorbed nitrate.
Torres-Dorante et al. (2008, 2009) observed that Mg–Al-LDH (4.5:1) The SEM micrograph of Mg–Al-LDH (4:1) showed a star-like pat-
had high nitrate adsorption capacity and selectivity in simulated soil tern resulting from well dispersion of plate-like particles. The XRD
solutions as well as soils. Although various LDHs have been investi- pattern exhibited sharp and symmetric reflections at 2θ = 10.98
gated for nitrate adsorption from aqueous solutions (Islam and (7.823 Ǻ) and 2θ = 22.06 (3.940 Ǻ) corresponding to the (003) and
Patel, 2009, 2010, 2011), the adsorption characteristics of the LDHs (006) diffraction planes which is characteristic of layered structures.
for nitrate compared with multivalent anions found in soil solutions The specific surface area of the studied LDH was 64.4 m 2g −1. The
have been less studied. PZC was measured about 12.5 and the anion exchange capacities
Preliminary experiments on four different types of LDHs including determined as 1.61 mmol g −1.
Mg–Al and Mg–Fe in 3:1 and 4:1 ratios of M 2+/M 3+ showed higher
nitrate adsorption capacity and selectivity of Mg–Al (4:1) compared 3.2. Adsorption kinetics
to other LDHs (Halajnia et al., 2012). Therefore, this LDH was selected
for further experiments. The objective of the present study was to For successful application of LDHs for nitrate removal, it is impor-
compare the adsorption characteristics of nitrate on Mg–Al LDH tant to achieve high adsorption capacity for nitrate in a short time.
(4:1) with those of sulfate and phosphate as common competing According to Fig. 1, the rate of anion removal by the LDH was sharp
anions in soil solution. at initial times followed by a long slow decline, as reported by Goh
et al. (2008). About 72, 54 and 58% of adsorption occured at the first
5 min for nitrate, phosphate and sulfate, respectively. The adsorption
2. Materials and methods equilibrium was fully established after 120 min for nitrate. For phos-
phate and sulfate, adsorption continued over 120 min (Fig. 1). The
All the chemicals used were of analytical grade and obtained from results showed that adsorption of nitrate was higher and faster than
Merck manufacture. Chloride form of Mg–Al with 4:1 ratio of M 2+: that of sulfate and phosphate. Islam and Patel (2009) showed that
M 3+ was synthesized using co-precipitation method, described in about 40% of nitrate adsorption on Mg–Al-LDH (3:1) occurred within
the previous paper (Halajnia et al., 2012). the first 5 min. Peng et al. (2009) reported that about 93% of adsorp-
The adsorption kinetics of nitrate, phosphate and sulfate on the tion of phosphate by Mg–Al-LDH (2:1) was obtained within 5 min.
selected LDH were investigated through batch experiments at initial Tsujimura et al. (2007) also observed adsorption of more than 95%
concentration of 5 mmol L −1 these anions as KNO3, K2HPO4 and on Mg–Al-LDH (2:1) for sulfate during the same adsorption time.
K2SO4, respectively for 5 to 120 min (solid-to-solution ratio = The percentage removal depends on the type and initial concentra-
5 g L −1, I = 0.03 mol L −1 KCl, pH = 7). The pH adjustments were tion of anion as well as LDH characteristics such as LDH type, calcina-
made by 0.1 mol L −1 HCl and 0.1 mol L −1 KOH. Adsorption iso- tion, crystallization of LDH, molar ratio of M 2+/M 3+and method of
therms for each anion were investigated at initial concentrations of synthesis (Das et al., 2006; Goh et al., 2008; Grover et al., 2010). Das
0.2 to 5 mmol L −1 and at three temperatures of 10, 25 and 50 °C et al. (2006) reported that the percentage of removal decreased
(solid-to-solution ratio = 5 g L −1, I = 0.03 mol L −1 KCl, pH = 7 with increasing initial phosphate concentration and increased with
and shaking time = 120 min). After ion adsorption from decreasing Mg 2+/Al 3+ molar ratio. In these studies, besides the dif-
5 mmol L −1 solutions at 25 °C, desorption experiments were carried ference in the initial concentrations, higher absorption rate of sulfate
out by replacing the 15 mL equilibrium solutions with the same vol- and phosphate within the first 5 min could be due to the lower
ume of 0.03 mol L −1 KCl and shaking for 120 min. Nitrate, phosphate Mg 2+/Al 3+ molar ratios. Also, in the present research, nitrate adsorp-
and sulfate concentrations in the extracts were measured by tion by Mg–Al (4:1) was higher than by the LDH with lower Mg 2+/
UV-spectrophotometeric method (Edwards et al., 2001), ascorbic Al 3+ molar ratio (3:1) reported by Islam and Patel (2009). This may
acid method (Murphy and Riley, 1962) and turbidimetric method be due to the higher nitrate adsorption capacity and selectivity of
(EPA, 1978), respectively. Moreover, the FTIR spectra of the LDHs Mg–Al (4:1) compared to Mg–Al (3:1), obtained from preliminary ex-
after adsorption of nitrate, phosphate and sulfate from 5 mmol L −1 periments (Halajnia et al., 2012).
solutions, were recorded using a Bruker Tensor 27 spectrometer. All The adsorption rate constant (k) and anion adsorbed (mmol g −1)
experiments were replicated twice. at equilibrium time for nitrate, phosphate and sulfate adsorption
In the isotherm experiments, the amount of the anion adsorbed
per unit mass of LDH in mmol g −1 was calculated by the following
0.7
equation:
0.6
Ion adsorbed (mmol g-1)

qe ¼ ðCi –Ce Þ  V=m 0.5 Nitrate


Phosphate
0.4
Sulfate
where Ci and Ce are initial and equilibrium concentrations of anion 0.3
(mmol L−1), m is the mass of LDH (g), and V is the volume of solution (L).
In the kinetic experiments, the amount of the anion adsorbed at 0.2
time t (mmol g −1) was obtained by the following equation:
0.1

0
qt ¼ ðCi –Ct Þ  V=m 0 10 20 30 40 50 60 70 80 90 100 110 120 130
Time (min)
where Ci and Ct are initial and at time t concentrations of anion Fig. 1. Nitrate, phosphate and sulfate adsorption on the LDH as a function of time
(mmol L−1), m is the mass of LDH (g), and V is the volume of solution (anion concentration = 5 mmol L−1, pH = 7, solid-to-solution ratio = 5 g L−1, I =
(L). 0.03 mol L−1 KCl at 25 °C).
A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312 307

Table 1
Mathematical forms and parameters of the kinetic models.

Kinetic model Linear form of model Parameters

Pseudo-first order log (qe − qt) = log (qe) − k1(t/2.303) t = time (min)
Pseudo-second order t/qt = 1/(k2qe2) + t/qe qe = ion adsorbed (mmol g−1) at equilibrium time
Intraparticle diffusion qt = k3t0.5 qt = ion adsorbed (mmol g−1) at time t
k = adsorption rate constant

were determined from the slope and intercept of linear plots of the high R 2 and low RMSE values, the good agreement of kinetics
pseudo-first order, pseudo second order and intraparticle diffusion data with the intraparticle diffusion model indicated that the
models (Chabani et al., 2006; Islam and Patel, 2010). Table 1 shows intraparticles diffusion may be considered as a rate limiting step in
the mathematical forms and parameters of these models. adsorption process. Nevertheless, the y-intercept in the model repre-
The root mean square error (RMSE) for the straight line fit of each sented that the boundary layer effect was another rate limiting factor
model was calculated as follows (Zhang and Selim, 2005). (Chatterjee and Woo, 2009; Demiral and Gunduzog˘lu, 2010). Results
showed that the rate of intraparticle diffusion (k3) for nitrate was
h   i0:5
 2 higher than that of sulfate and phosphate (Table 2). Since ion diffu-
RMSE ¼ ∑ qt –qt =ðn–mÞ
sion in layered materials like LDHs is highly dependent on ion size
and layer network (Tezuka et al., 2004, 2005), the higher intraparticle
where qt and qt* are the amount of nitrate adsorbed (mmol g −1) at adsorption rate constant (k3) for nitrate indicates the importance of
time t (min), obtained from laboratory experiments and predicted diffusion into interlayer spaces in nitrate adsorption. The values of
by the models, respectively. The n value is the number of measure- y-intercept give an idea about the boundary layer thickness and
ments and m is the number of fitted parameters. increase with increasing ion concentration around the particles. Due
The determination coefficients (R 2) and RMSE for estimating the to the greater y-intercept values, the effect of boundary layer on
goodness-of-fit of the models to the experimental data are given in nitrate adsorption was higher than that of phosphate and sulfate. In
Table 2. other words, the greater y-intercept values as a result of the increase
As evidenced by the higher R 2 and lower RMSE values (Table 2), in thickness of the boundary layer indicate an increased chance of
the experimental data of nitrate, phosphate and sulfate adsorption internal mass transfer (Santhi et al., 2010). Therefore, it seems that
on the LDH were better fitted to the pseudo-second order model. In for nitrate adsorption on the synthesized LDH, diffusion throughout
addition, for all three anions the calculated ion adsorbed at equilibri- the internal spaces was more important than surface adsorption.
um time (qe) from the pseudo second order model was more close to
the experimental data. The experimental qe values for nitrate, phos- 3.3. Thermodynamic parameters for adsorption of anions
phate and sulfate were obtained, 0.590, 0.282 and 0.122 mmol g −1,
respectively. Among the kinetics models, pseudo first order and pseu- Soil temperature is influenced by variations in air temperature and
do second order have been widely used for oxyanion adsorption by changes daily and annually. Hence, before applying LDHs to the soils,
LDHs (Goh et al., 2008). These common models are used in order to it is important to understand the effect of temperature on ion adsorp-
determine the rate constants of adsorption and to find out informa- tion from soil solution. Nitrate, phosphate and sulfate adsorption
tion about the adsorption mechanisms. isotherms at temperatures of 10, 25 and 50 °C are shown in Figs. 2,
The good fit of the experimental data with the pseudo first order 3 and 4. Phosphate and sulfate adsorption increased with increasing
model was reported by Islam and Patel (2009, 2010, 2011) for nitrate temperature but nitrate adsorption decreased. These results indicate
adsorption on Mg–Al-LDH (3:1), Zn–Al-LDH (3:1) and Ca–Al-LDH the exothermic nature of nitrate and endothermic nature of sulfate
(3:1) and by Das et al. (2006) for phosphate adsorption on Mg– and phosphate adsorption by the synthesized LDH. Since the highest
Al-LDH (2:1). Cheng et al. (2009), Hosni and Srasra (2010) and Cai nitrate leaching normally occurs in the cold, rainy and non-growing
et al. (2012) reported the better performance of pseudo second seasons, particularly from late autumn until early spring, increasing
order to describe phosphate adsorption on Zn–Al-LDH (2:1), Mg– nitrate adsorption with decreasing temperature is considered as a
Al-LDH (3:1) and Mg–Al-LDH (2:1), respectively. Some possible good advantage of nitrate adsorption processes by this LDH. On the
causes of discrepancies between kinetic models in literature includ- other hand, competition of nitrate with phosphate and sulfate is
ing different experimental conditions, random errors in data acquisi- reduced due to decreasing adsorption of these anions in this period.
tion and artifacts in the sorption experiments and pre-qualifying a Thermodynamic parameters for adsorption of nitrate, phosphate
few possible kinetic models were discussed by Goh et al. (2008). and sulfate on the synthesized LDH were determined. The Gibbs
Several steps are involved in the adsorption process, i.e. (a) mass free energy (ΔGro) for the adsorption process was calculated as
transfer from the bulk solution to the particle surface, (b) film diffu- follows (Chatterjee and Woo, 2009):
sion through the boundary layer surrounding the particle, (c) internal
o
diffusion within the particle and (d) surface adsorption (Singh and ΔGr ¼ –RT lnðKL Þ
Pant, 2006). In the present study conditions, mass transfer from the
bulk solution cannot be the rate limiting step because of shaking where R, T, and KL are gas constant, temperature in Kelvin, and ad-
fast enough the suspensions. For all three anions, as evidenced by sorption equilibrium constant of Langmuir isotherm, respectively. In

Table 2
Kinetic parameters for nitrate, phosphate and sulfate adsorption by the LDH.

Pseudo-first order model Pseudo-second order model Intraparticle diffusion model


−1 −1 −1 −1 −1
Ions qe (mmol g ) k1 (min ) R 2
RMSE qe (mmol g ) k2 (g mmol min ) R 2
RMSE Intercept k3 (mmol g−1.min−0.5) R2 RMSE

Nitrate 0.256 0.088 0.997 0.445 0.619 0.623 0.999 0.010 0.332 0.046 0.960 0.013
Phosphate 0.141 0.032 0.983 0.201 0.286 0.593 0.997 0.014 0.116 0.020 0.946 0.012
Sulfate 0.078 0.048 0.985 0.064 0.130 1.056 0.991 0.012 0.048 0.009 0.985 0.003
308 A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312

0.7 0.18
10 oC 25 oC 50 oC 10 oC 25 oC 50 oC

Sulfate adsorbed (mmol g-1)


Nitrate adsorbed (mmol g-1)

0.6 0.16
0.14
0.5
0.12
0.4 0.1
0.3 0.08

0.2 0.06
0.04
0.1
0.02
0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Ce (mmol L-1) Ce (mmol L-1)

Fig. 2. Equilibrium isotherms of nitrate adsorption on the LDH as a function of temper- Fig. 4. Equilibrium isotherms of sulfate adsorption on the LDH as a function of temper-
ature (°C) (pH = 7, solid-to-solution ratio = 5 g L−1, I = 0.03 mol L−1 KCl). ature (°C) (pH = 7, solid-to-solution ratio = 5 g L−1, I = 0.03 mol L−1 KCl).

order to obtain a correct value of ΔGr o, the KL values in mol L −1 the phosphate adsorption by calcined Mg–Al-LDH (2:1) decreased
became dimensionless by multiplying it by 55.5 (number of moles of with increasing temperature. Comparable results were observed for
water per liter of solution) (Bhatnagar and Minocha, 2010; Milonjic, phosphate adsorption on Mg–Mn-LDH (3:1) by Chitrakar et al.
2007). (2005) and on Zn–Al-LDH (2:1) by Cheng et al. (2009). Few studies
The enthalpy (ΔHr o) and entropy (ΔSr o) of adsorption were calcu- have been performed on sulfate adsorption on LDHs. Tsujimura et al.
lated using the slope and intercept of van't Hoff linear equation, (2007) reported that sulfate removal by Mg–Al-LDH was independent
respectively as follows (Chabani et al., 2006): of temperature. It seems that structural differences, molar ratio of
M 2+/M 3+, metal cations in the LDH and experimental conditions
o o
lnðKL Þ ¼ –ΔHr =RT þ ΔSr =R: have strong influence on the adsorption process. The electrostatic
adsorption is usually exothermic in nature. The endothermic nature
Thermodynamics parameters for the anions are reported in could be due to chemical interaction, increasing the reactivity of the
Table 3. For all three anions the negative values of ΔGr o indicated surface sites and the rate of intraparticle diffusion of sorbate ions
the spontaneous nature of adsorption process. The higher negative into the pores of adsorbent at higher temperature (Chabani et al.,
values of ΔGr o for phosphate and sulfate adsorption at higher temper- 2006; Chitrakar et al., 2005; Das et al., 2006; Goh et al., 2008). The
atures showed that adsorption of these anions were more energeti- values of ΔSr° were positive for studied anions. The positive values of
cally favorable than that of nitrate due to higher ΔSr o. For nitrate ΔSr° indicate the increased randomness at the solid–solution interface
ΔSr o has a low value, therefore ΔGr o did not change significantly. during adsorption. Positive values of ΔSr° have been widely reported
The negative values of ΔHr o confirmed the exothermic nature of ni- for anion adsorption on LDHs (Das et al., 2006, 2003; Islam and
trate adsorption. Whereas, the positive values of ΔHro for sulfate and Patel, 2009, 2010, 2011).
phosphate indicated that the adsorption process was endothermic
in nature. These findings for nitrate are in agreement with the
findings of Khataee and Khani (2009), Chabani et al. (2006) and 3.4. Adsorption–desorption isotherms
Chatterjee and Woo (2009) for nitrate adsorption on granular acti-
vated carbon, anionic exchange resin and chitosan, respectively, but Adsorption–desorption of anions on LDHs influence their removal,
in disagreement with the reports of Islam and Patel (2009, 2010, transportation and bioavailability in aqueous solutions and soil envi-
2011) for nitrate adsorption by Mg–Al, Zn–Al and Ca–Al-LDH (3:1). ronments. Nitrate, phosphate and sulfate adsorption–desorption
Tezuka et al. (2005) also reported the endothermic nature of nitrate isotherm at 25 °C are shown in Fig. 5. The apparent distribution
adsorption by Ni–Fe LDH. Although, Das et al. (2006) showed that

0.7
0.45 Nitrate-Adsorption
10 oC 25 oC 50 oC
Phosphate adsorbed (mmol g-1)

0.6 Nitrate-Desorption
Ion adsorbed (mmol g-1)

0.4
phosphate-Adsorption
0.35 0.5 phosphate-Desorption
0.3 Sulfate-Adsorption
0.4
0.25 Sulfate-Desorption
0.3
0.2
0.15 0.2
0.1
0.1
0.05
0
0 0 1 2 3 4 5
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
C (mmol L-1)
Ce (mmol L-1)
Fig. 5. Equilibrium adsorption–desorption isotherms of sulfate, phosphate and nitrate
Fig. 3. Equilibrium isotherms of phosphate adsorption on the LDH as a function of tem- adsorption on the LDH (pH = 7, solid-to-solution ratio = 5 g L−1, I = 0.03 mol L−1
perature (°C) (pH = 7, solid-to-solution ratio = 5 g L−1, I = 0.03 mol L−1 KCl). KCl at 25 °C).
A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312 309

Table 3 where Kd (L g −1) and m are the apparent distribution coefficient and
The values of ΔGr 0 and ΔHr 0 for adsorption of sulfate, phosphate and nitrate by the LDH. total number of desorption steps, respectively. The index is zero for
Ion ΔHr o (kJ mol−1) ΔSr o (J mol−1 K−1) ΔGr o (kJ mol−1) completely reversible reactions and approaches one as the reaction
tends toward complete irreversibility. The HI value for phosphate
10 °C 25 °C 50 °C
(0.74) was markedly higher than that of sulfate (0.14) and nitrate
Nitrate −11.87 38.63 −22.91 −23.20 −24.43
(0.02).
Phosphate 39.52 236.07 −26.08 −32.96 −35.79
Sulfate 1.71 105.17 −28.05 −29.65 −32.26 Kuzawa et al. (2006) observed low phosphate desorption efficien-
cies for Mg–Al-LDH (2:1) by 30% (w/v) NaCl solution. Nevertheless,
an effective desorption was achieved with alkaline NaCl solutions.
coefficient (Kd) is a factor related to distribution of ion between solid Lv et al. (2008) also reported 12% phosphate desorption efficiency
and aqueous phases and was calculated by the following expression for Mg–Al-LDH (2:1) using a 1 M NaCl solution. They showed that
(Torres-Dorante et al., 2008): phosphate desorption in presence of different anions decreased in
order of CO23− > OH − > Cl −. This may be contributed to different
Kd ¼ qe =Ce affinities of ions towards LDHs.
Adsorption–desorption processes are usually described through
where qe is the amount of ion adsorbed per unit mass of the LDH and isotherms that represent the adsorbate distribution between liquid
Ce is the ion concentration per unit volume of solution. and solid phases as a function of equilibrium concentration.
For all three anions decreasing in apparent distribution coefficient Freundlich, Langmuir and Dubinin–Randushkevich isotherms were
values with increasing ion concentration in equilibrium solution used to describe the adsorption–desorption characteristics of studied
(Table 4) indicated that the LDH affinity for the studied anions was anions (Chabani et al., 2006; Das et al., 2006; Foo and Hameed, 2010).
greater in lower concentrations, probably due to the saturation of The mathematical forms of the models and their parameters are given
adsorption sites. However, the estimated range of Kd values for ni- in Table 5. The values of parameters for each model are shown in
trate was less than two other anions. In other words, the Kd values Table 6.
were almost independent of the surface coverage, probably due to The higher values of R 2 indicated that the adsorption data were
the nitrate adsorption into the interlayer spaces. The greater Kd values better fit to the Freundlich model for nitrate and to the Langmuir
for phosphate at initial concentrations of 0.2, 0.5 and 1 mmol L −1 model for phosphate and sulfate. Comparable results were obtained
indicated the remarkable affinity of phosphate for the synthesized by Cheng et al. (2009) and Hosni and Srasra (2010) for phosphate
LDH at lower concentrations. and by Tezuka et al. (2005) for nitrate. However, the results were in
In addition to adsorption capacity and selectivity, the reversibility disagreement with the findings of Islam and Patel (2009, 2010,
of adsorption is important for using LDH as a nitrate exchanger in soil. 2011) and Hosni and Srasra (2008) for nitrate and with the findings
For this reason, nitrate desorption experiments were performed. For of Peng et al. (2009) and Chitrakar et al. (2005) for phosphate. The
all three anions, a gap was observed between the adsorption and experimental adsorption data for three anions were best described
desorption branches of the isotherms (Fig. 5). Possible explanations by Dubinin–Radushkevich model. The goodness of fit to this model
for this hysteresis include (1) irreversible chemical binding, (2) slow also was reported by Islam and Patel (2009, 2010, 2011) for nitrate.
rates of desorption, and (3) microporus entrapment of ions (Chefetz The experimental desorption data were best described by the
et al., 2004; Huang et al., 2003). Freundlich model for phosphate and sulfate and by the Langmuir
The percentage of nitrate, phosphate and sulfate desorbed from model for nitrate.
the LDH after five cycles were 66.3, 15.7 and 58.8%, respectively. Apparently, the goodness-of-fit to the Langmuir equation for
According to Cox et al. (1997) desorption hysteresis coefficient (H) phosphate and sulfate indicates monolayer adsorption of ions on
was calculated by the following equation: the LDH surface. Whereas, the Freundlich model for nitrate repre-
sents that the adsorption is not limited by monolayer coverage. This
H ¼ ðð1=ndes Þ=ð1=nad ÞÞ  100 empirical model is widely used for multilayer adsorption on hetero-
geneous surfaces (Foo and Hameed, 2010). It is better to say that
where 1/ndes and 1/nad are exponents of Freundlich equation for probably adsorption of nitrate by the synthesized LDH is not limited
desorption and adsorption branches of isotherm, respectively. The to the external surfaces. Hence, higher adsorption capacity of this
lower H values indicate more difficult desorption to the solution. LDH for nitrate might be attributed to higher nitrate adsorption
The H values for nitrate, phosphate and sulfate were 97.5, 22.5 and onto internal surfaces. A few studies have reported that the high
79.0%, respectively. The results showed that nitrate adsorption was capacity of some LDHs for nitrate is strongly influenced by the NO3−
more reversible than phosphate and sulfate. Phosphate adsorption ion-sieve property due to the spatial limits between the LDH brucite
was less reversible probably due to specific adsorption. Furthermore, layers (Sasai et al., 2012; Tezuka et al., 2004). It seems that even the
the same results can be also obtained by calculating a hysteresis index orientation of the interlayer ions in LDHs can affect the absorption
based on the apparent distribution coefficients as follows (Laird et al., capacity. The influence of the layer charge density on the orientation
1994): of the interlayer nitrate in Mg–AL-LDHs with the Mg:Al ratios of 2, 3
and 4 were investigated by Wang and Wang (2007). They reported
  different orientation for studied LDHs and a horizontal orientation
i i−1
HI ¼ ∑K d −K d =m
for the interlayer nitrate in Mg–Al-LDH (4:1). According to Table 6,
the value of constant related to binding energy (KL) for phosphate
adsorption (10.731 L mmol −1) was greater than that of sulfate
Table 4 (3.021 L mmol −1) and nitrate (0.210 L mmol −1) that can be due to
Apparent distribution coefficient values (L g−1) for nitrate, phosphate and sulfate.
the formation of stronger adsorption bonds. Whereas, the value of
Initial C (mmol L−1) Nitrate Phosphate Sulfate maximum adsorption capacity (b) for nitrate (1.905 mmol g −1)
0.2 0.42 12.61 0.51 was considerably greater than that of sulfate (0.278 mmol g −1) and
0.5 0.42 3.34 0.24 phosphate (0.126 mmol g −1). The same trend of values were found
1 0.42 0.78 0.12 for the desorption isotherms. According to the chemical analysis,
3 0.32 0.14 0.04 the empirical formula of the synthesized LDH should be Mg0.802
5 0.28 0.08 0.03
Al0.198(OH)2 Cl0.198 0.57 H2O. Therefore, the theoretical anion exchange
310 A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312

Table 5
Mathematical forms of the studied isotherms and their parameters.

Adsorption isotherm model Linear form of the model Parameters

Freundlich log qe = log KF + (1/n) log Ce qe = ion adsorbed per unit mass of LDH (mmol g−1)
Langmuir Ce/qe = Ce/b + 1/(KLb) Ce = ion equilibrium concentration (mmol L−1)
Dubinin–Randushkevich ln qe = ln qm − K 2 kF = experimental constant
n = experimental constant
kL = constant related to binding energy
b = maximum adsorption capacity for monolayer coverage
qm = theoretical adsorption capacity

Table 6
The parameters of Freundlich, Langmuir and Dubinin–Randushkevich isotherms for absorption and desorption of nitrate, phosphate and sulfate by the LDH.

Freundlich model Langmuir model Dubinin–Randushkevich model

Ions KF n R2 RMSE b KL R2 RMSE qm K R2 E RMSE


(mmol(1 − 1/n) g−1 L1/n) (mmol g−1) (Lmmol−1) (mmol g−1) (mol2 kJ−2) (kJ mol−1)

Adsorption
Nitrate 0.318 1.116 0.992 0.024 1.905 0.210 0.734 0.011 4.979 0.009 0.992 7.344 0.012
Phosphate 0.216 3.704 0.974 0.022 0.278 10.731 0.997 0.032 0.459 0.002 0.996 15.811 0.038
Sulfate 0.080 3.106 0.979 0.007 0.126 3.021 0.995 0.009 0.231 0.003 0.995 12.910 0.016

Desorption
Nitrate 0.486 1.175 0.983 0.019 1.971 0.320 0.778 0.015 7.590 0.009 0.988 7.293 0.016
Phosphate 0.279 16.474 0.927 0.006 0.277 75.669 1.000 0.005 0.332 0.001 0.946 28.868 0.005
Sulfate 0.102 3.931 0.987 0.003 0.113 9.380 0.997 0.006 0.214 0.003 0.994 14.030 0.002

capacity which is calculated from the structural formula is two anions. The E value from this model gives an idea about the type
2.61 mmol g−1. Although studied LDH displayed high affinity to ni- of adsorption. The E values less than 8 kJ mol −1 characterize a physi-
trate, but the maximum adsorption capacity of nitrate calculated by cal adsorption type and 8 b E b 16 kJ mol −1 indicates the chemical
Langmuir model was less than the theoretical AEC. In addition to the ion exchange (Chabani et al., 2006; Itodo and Itodo, 2010). The E
type of LDH and anion, adsorption capacity is strongly influenced by value for nitrate was obtained less than 8 kJ mol −1 which corre-
ionic strength, pH, adsorbent dose and the presence of other ions sponds to physical adsorption of nitrate. Whereas, the greater values
(Goh et al., 2008). of E for phosphate and sulfate indicate chemical adsorption of these
The final equilibrium pH values after nitrate, phosphate and sul- anions on the LDH. The E values from this model for sulfate and
fate adsorption changed between 8.7 and 8.9. This is attributed to especially for phosphate were higher on the desorption isotherm
high buffering capacity or partial dissolution of the LDH (Cheng et than on the adsorption isotherm indicating more difficult desorption
al., 2009; Yang et al., 2012). of these ions to the solution.
Experimental data were also fit well with Dubinin–Radushkevich The K constant related to the adsorption energy was greater for
model. The theoretical adsorption capacity of Dubinin–Randushkevich nitrate than sulfate and phosphate. It seems that due to the physical
isotherm (qm) for nitrate was remarkable higher than that of other nature of nitrate adsorption, ion-sieve property of studied LDH is the

Fig. 6. The FTIR spectra of the LDH before and after nitrate, phosphate and sulfate adsorption.
A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312 311

Fig. 7. XRD patterns of the LDH before and after nitrate, phosphate and sulfate adsorption.

main reason for higher nitrate adsorption capacity. Chitrakar et al. appropriate interlayer distance and the proper orientation of nitrate
(2005) suggested that a chemical interaction involves the adsorption ions with respect to the LDH layers.
of phosphate on the LDH. According to the Raman spectroscopy stud-
ies, Frost et al. (2005) proposed two bonding types for XO4 anions 4. Conclusion
including hydrated and hydroxyl surface bonded. Although, Islam
and Patel (2009) reported a chemical ion exchange process for nitrate The synthesized Mg–Al-LDH (4:1) exhibited higher adsorption
adsorption on Mg–Al-LDH (3:1). The physical nature of nitrate rate and adsorption capacity to nitrate compared with phosphate
adsorption has been reported by Chabani et al. (2006) and Milmile and sulfate. This LDH can be used as an effective adsorbent for removal
et al. (2011) on anionic exchange resin and by Bhatnagar et al. of nitrate from soil solution. Calculated thermodynamic parameters
(2010) on nano-alumina. showed that the adsorption process was endothermic for phosphate
The greater affinity of the synthesized LDH toward nitrate than and sulfate and exothermic for nitrate. In this regard, as a result of
for phosphate and sulfate was also confirmed by FTIR spectroscopy. increasing nitrate adsorption with decreasing temperature, the syn-
The FTIR spectra of studied LDH before and after nitrate, phosphate thesized LDH can efficiently reduce the risk of nitrate leaching during
and sulfate adsorption were shown in Fig. 6. The absorption bands the cold and nongrowing seasons. Nitrate showed higher adsorption
around 3500 cm−1 and 570–670 cm−1 belong to OH stretching vibra- rate compared to phosphate and sulfate. Adsorption of nitrate on the
tions (from structural hydroxyl groups and interlayer water molecules) synthesized LDH was more reversible than that of phosphate and
and coupled MgO6 and AlO6 octahedral, respectively. The LDH structure sulfate which is important for nitrate availability to plants.
is characterized by these bands (Olfs et al., 2009). The absorption bands
at 1700–1610 cm−1 were assigned as the O\H bending mode of water
Acknowledgements
molecules (Chuang et al., 2008). The absorption band at 1383 cm−1
indicated the presence of nitrate ions intercalated in the interlayer
We are thankful for the financial support of Tabriz University and
space of the LDH (Goh et al., 2010; Olfs et al., 2009; Tezuka et al.,
we also thank to Ferdowsi University of Mashhad for providing facil-
2004). A sharp absorption band at 1383 cm−1 indicated the large
ities to carrying out this work.
amounts of intercalated nitrate ions. The FTIR spectra did not show
such a strong band for phosphate and sulfate. The P\O stretching vibra-
tion leads to a distinct band at 1050 cm−1 (Cheng et al., 2010; He et al., References
2010; Hosni and Srasra, 2010). A band at 1120 cm−1 is also attributed Bhatnagar, A., Minocha, A.K., 2010. Biosorption optimization of nickel removal from
to the coverage of sulfate ions on the synthesized LDH. water using Punica granatum peel waste. Colloids and Surfaces. B, Biointerfaces
Fig. 7 shows the XRD patterns of Mg–Al-LDH before and after 76, 544–548.
Bhatnagar, A., Kumar, E., Sillanpää, M., 2010. Nitrate removal from water by nano-alumina:
nitrate, phosphate and sulfate exchange. The results showed that characterization and sorption studies. Chemical Engineering Journal 163, 317–323.
the basal d-spacing of Mg–Al-LDH was significantly reduced after Cai, P., Zheng, H., Wang, C., Ma, H., Hu, J., Pu, Y., Liang, P., 2012. Competitive adsorption
nitrate exchange. Although the d-spacing of LDHs strongly depends characteristics of fluoride and phosphate on calcined Mg–Al–CO3 layered double
hydroxides. Journal of Hazardous Materials 213–214, 100–108.
on the charge, size and consequently the extent of hydration of Chabani, M., Amrane, A., Bensmaili, A., 2006. Kinetic modelling of the adsorption of
interlayer anions but Del Arco et al. (2000), Wang and Wang (2007) nitrates by ion exchange resin. Chemical Engineering Journal 125, 111–117.
and Wang et al. (2009) reported that the reorientation of nitrates Chatterjee, S., Woo, S.H., 2009. The removal of nitrate from aqueous solutions by chitosan
hydrogel beads. Journal of Hazardous Materials 164, 1012–1018.
intercalated into the interlayer region of Mg–Al-LDHs as a conse-
Chefetz, B., Bilkis, Y.I., Polubesova, T., 2004. Sorption–desorption behavior of triazine and
quence of changing layer charge density is another factor affecting phenylurea herbicides in Kishon river sediments. Water Research 38, 4383–4394.
the basal spacing and one of the major factors governing the nitrate Cheng, X., Huang, X., Wang, X., Zhao, B., Chen, A., Sun, D., 2009. Phosphate adsorption
from sewage sludge filtrate using zinc–aluminum layered double hydroxides. Journal
adsorption capacity and selectivity. For the Mg–Al-LDH (2:1) with
of Hazardous Materials 169, 958–964.
high layer charge density, a large basal spacing is attributed to the Cheng, X., Huang, X., Wang, X., Sun, D., 2010. Influence of calcination on the adsorptive
vertical or tilted arrangement of nitrate ions but a flat orientation of removal of phosphate by Zn–Al layered double hydroxides from excess sludge
nitrate ions parallel to the hydroxide sheets of the Mg–Al-LDH (4:1) liquor. Journal of Hazardous Materials 177, 516–523.
Chitrakar, R., Tezuka, S., Sonoda, A., Sakane, K., Ooi, K., Hirotsu, T., 2005. Adsorption of
results in small interlayer space. It seems that the high nitrate adsorp- phosphate from seawater on calcined MgMn-layered double hydroxides. Journal of
tion capacity and selectivity of the Mg–Al-LDH (4:1) is due to the Colloid and Interface Science 290, 45–51.
312 A. Halajnia et al. / Applied Clay Science 80–81 (2013) 305–312

Chuang, Y.H., Tzou, Y.M., Wang, M.K., Liu, C.H., Chiang, P.N., 2008. Removal of Komarneni, S., Newalkar, B.L., Li, D., Gheyi, T., 2003. Anionic clays as potential slow-
2-chlorophenol from aqueous solution by Mg/Al layered double hydroxide (LDH) release fertilizers: nitrate ion exchange. Journal of Porous Materials 10, 243–248.
and modified LDH. Industrial and Engineering Chemistry Research 47, 3813–3819. Kuzawa, K., Jung, Y.J., Kiso, Y., Yamada, T., Nagai, M., Lee, T.G., 2006. Phosphate removal
Cox, L., Koskinen, W.C., Yen, P.Y., 1997. Sorption–desorption of imidacloprid and its and recovery with a synthetic hydrotalcite as an adsorbent. Chemosphere 62,
metabolites in soils. Journal of Agricultural and Food Chemistry 45, 1468–1472. 45–52.
Das, D.P., Das, J., Parida, K., 2003. Physicochemical characterization and adsorption be- Laird, D.A., Yen, P.Y., Koskinen, W.C., Steinheimer, T.R., Dowdy, R.H., 1994. Sorption of
havior of calcined Zn/Al hydrotalcite-like compound (HTlc) towards removal of atrazine on soil clay components. Environmental Science and Technology 28,
fluoride from aqueous solution. Journal of Colloid and Interface Science 261, 1054–1061.
213–220. Lv, L., Sun, P., Wang, Y., Du, H., Gu, T., 2008. Phosphate removal and recovery with
Das, J., Patra, B.S., Baliarsingh, N., Parida, K.M., 2006. Adsorption of phosphate by calcined layered double hydroxides as an adsorbent. Phosphorus, Sulfur, and Silicon
layered double hydroxides in aqueous solutions. Applied Clay Science 32, 252–260. and the Related Elements 183, 519–526.
Del Arco, M., GutieHrrez, S., MartmHn, C., Rives, V., Rocha, J., 2000. Effect of the Mg:Al Milmile, S.N., Pande, J.V., Karmakar, S., Bansiwal, A., Chakrabarti, T., Biniwale, R.B., 2011.
ratio on borate (or silicate)/nitrate exchange in hydrotalcite. Journal of Solid State Equilibrium isotherm and kinetic modeling of the adsorption of nitrates by anion
Chemistry 151, 272–280. exchange Indion NSSR resin. Desalination 276, 38–44.
Demiral, H., Gunduzog˘lu, G., 2010. Removal of nitrate from aqueous solutions by activated Milonjic, S.K., 2007. A consideration of the correct calculation of thermodynamic
carbon prepared from sugar beet bagasse. Bioresource Technology 101, 1675–1680. parameters of adsorption. Journal of the Serbian Chemical Society 72, 1363–1367.
Edwards, A.C., Hooda, S.P., Cook, Y., 2001. Determination of nitrate in water containing Murphy, J., Riley, J.P., 1962. A modified single solution method for the determination of
dissolved organic carbon by ultraviolet spectroscopy. International Journal of phosphate in natural waters. Analytica Chimica Acta 27, 31–36.
Environmental Analytical Chemistry 80, 49–59. Olfs, H.W., Torres-Dorante, L.O., Eckelt, R., Kosslick, H., 2009. Comparison of different
EPA Method 375.4., 1978. Sulfate (Turbidimetric). Methods and guidance for analysis synthesis routes for Mg–Al layered double hydroxides (LDH): characterization of
of water, EPA 821-C-97-001. United States Environmental Protection Agency, the structural phases and anion exchange properties. Applied Clay Science 43,
Office of Water, Washington, DC. 459–464.
Foo, K.Y., Hameed, B.H., 2010. Insights into the modeling of adsorption isotherm sys- Peng, S., Lu, L., Wang, J., Han, L., Chen, T., Jiang, S., 2009. Study on the adsorption kinetics of
tems. Chemical Engineering Journal 156, 2–10. orthophosphate anions on layer double hydroxide. Chinese Journal of Geochemistry
Forano, C., Hibino, T., Leroux, F., Taviot-Gueho, C., 2006. Layered double hydroxides. In: 28, 184–187.
Bergaya, F., Theng, B.K.G., Lagaly, G. (Eds.), Handbook of Clay Science. Elsevier Ltd., Prakasa Rao, E.V.S., Puttanna, K., 2006. Strategies for combating nitrate pollution.
pp. 1021–1095. Current Science 91, 1335–1339.
Frost, R.L., Musumeci, A.W., Martens, W.N., Adebajo, M.O., Bouzaid, J., 2005. Raman Santhi, T., Manonmani, S., Smitha, T., 2010. Kinetics and isotherm studies on cationic
spectroscopy of hydrotalcites with sulphate, molybdate and chromate in the interlayer. dyes adsorption onto Annona Squmosa seed activated carbon. International Journal
Journal of Raman Spectroscopy 36, 925–931. of Engineering, Science and Technology 2, 287–295.
Génin, J.M.R., Refait, P., Bourrié, G., Abdelmoula, M., Trolard, F., 2001. Structure and Sasai, R., Norimatsu, W., Matsumoto, W., 2012. Nitrate-ion-selective exchange ability of
stability of the Fe(II)–Fe(III) green rust “fougerite” mineral and its potential for layered double hydroxide consisting of MgII and FeIII. Journal of Hazardous Materials
reducing pollutants in soil solutions. Applied Geochemistry 16, 559–570. 215–216, 311–314.
Goh, K.H., Lim, T.T., Dong, Z., 2008. Application of layered double hydroxides for removal Singh, T.S., Pant, K.K., 2006. Kinetics and mass transfer studies on the adsorption of
of oxyanions: a review. Water Research 42, 1343–1368. arsenic onto activated alumina and iron oxide impregnated activated alumina.
Goh, K.H., Lim, T.T., Banas, A., Dong, Z., 2010. Sorption characteristics and mechanisms of Water Quality Research Journal of Canada 4, 147–156.
oxyanions and oxyhalides having different molecular properties on Mg/Al layered Tezuka, S., Chitrakar, R., Sonoda, A., Ooi, Tomida, T., 2004. Studies on selective adsor-
double hydroxide nanoparticles. Journal of Hazardous Materials 179, 818–827. bents for oxo-anions. Nitrate ion-exchange properties of layered double hydrox-
Grover, K., Komarneni, S., Katsuki, H., 2010. Synthetic hydrotalcite-type and ides with different metal atoms. Green Chemistry 6, 104–109.
hydrocalumite-type layered double hydroxides for arsenate uptake. Applied Clay Tezuka, S., Chitrakar, R., Sonoda, A., Ooi, K., Tomida, T., 2005. Studies on selective adsor-
Science 48, 631–637. bents for oxo-anions. NO3− adsorptive properties of Ni–Fe layered double hydroxide
Halajnia, A., Oustan, S., Najafi, N., Khataee, A.R., Lakzian, A., 2012. The adsorption char- in seawater. Adsorption 11, 751–755.
acteristics of nitrate on Mg–Fe and Mg–Al layered double hydroxides in a simulated Torres-Dorante, L.O., Lammel, J., Kuhlmann, H., Witzke, T., Olfs, H.W., 2008. Capacity,
soil solution. Applied Clay Science 70, 28–36. selectivity, and reversibility for nitrate exchange of a layered double-hydroxide
He, H., Kang, H., Ma, S., Bai, Y., Yang, X., 2010. High adsorption selectivity of Zn–Al lay- (LDH) mineral in simulated soil solutions and in soil. Journal of Plant Nutrition
ered double hydroxides and the calcined materials toward phosphate. Journal of and Soil Science 171, 777–784.
Colloid and Interface Science 343, 225–231. Torres-Dorante, L.O., Lammel, J., Kuhlmann, H., 2009. Use of a layered double hydroxide
Hosni, K., Srasra, E., 2008. Nitrate adsorption from aqueous solution by MII–Al–CO3 lay- (LDH) to buffer nitrate in soil: long-term nitrate exchange properties under
ered double hydroxide. Inorganic Materials 44, 742–749. cropping and fallow conditions. Plant and Soil 315, 257–272.
Hosni, K., Srasra, E., 2010. Evaluation of phosphate removal from water by calcined- Tsujimura, A., Uchida, M., Okuwaki, A., 2007. Synthesis and sulfate ion-exchange prop-
LDH synthesized from the dolomite. Colloid Journal 72, 423–431. erties of a hydrotalcite-like compound intercalated by chloride ions. Journal of
Huang, W., Peng, P., Yu, Z., Fu, J., 2003. Effects of organic matter heterogeneity on Hazardous Materials 143, 582–586.
sorption and desorption of organic contaminants by soils and sediments. Applied Wang, S.L., Wang, P.C., 2007. In situ XRD and ATR-FTIR study on the molecular orienta-
Geochemistry 18, 955–972. tion of interlayer nitrate in Mg/Al-layered double hydroxides in water. Colloids and
IAEA (International Atomic Energy Agency), 2008. Guidelines on nitrogen manage- Surfaces A: Physicochemical and Engineering Aspects 292, 131–138.
ment in agricultural systems. Training course series, NO, 29. Wang, S.L., Liu, C.H., Wang, M.K., Chuang, Y.H., Chiang, P.N., 2009. Arsenate adsorption
Islam, M., Patel, R., 2009. Nitrate sorption by thermally activated Mg/Al chloride by Mg/Al–NO3 layered double hydroxides with varying the Mg/Al ratio. Applied
hydrotalcite-like compound. Journal of Hazardous Materials 169, 524–531. Clay Science 43, 79–85.
Islam, M., Patel, R., 2010. Synthesis and physicochemical characterization of Zn/Al Yang, Y., Gao, N., Chu, W., Zhang, Y., Ma, Y., 2012. Adsorption of perchlorate from aqueous
chloride layered double hydroxide and evaluation of its nitrate removal efficiency. solution by the calcination product of Mg/(Al–Fe) hydrotalcite-like compounds.
Desalination 256, 120–128. Journal of Hazardous Materials 209–210, 318–325.
Islam, M., Patel, R., 2011. Physicochemical characterization and adsorption behavior of You, Y., Vance, G.F., Zhao, H., 2001. Selenium adsorption on Mg–Al and Zn–Al layered
Ca/Al chloride hydrotalcite-like compound towards removal of nitrate. Journal of double hydroxides. Applied Clay Science 20, 13–25.
Hazardous Materials 90, 659–668. Zhang, H., Selim, H.M., 2005. Kinetics of Arsenate adsorption–desorption in Soils. Envi-
Itodo, A.U., Itodo, H.U., 2010. Sorption energies estimation using Dubinin-Radushkevich ronmental Science and Technology 39, 6101–6108.
and Temkin adsorption isotherms. Life Science Journal 7, 31–39.
Khataee, A.R., Khani, A., 2009. Modeling of nitrate adsorption on granular activated
carbon (GAC) using artificial neuron network (ANN). International Journal of
Chemical Reactor Engineering 7, 1–16.

You might also like