You are on page 1of 11

Environ Monit Assess (2013) 185:10365–10375

DOI 10.1007/s10661-013-3338-5

Removal of 8-quinolinecarboxylic acid pesticide


from aqueous solution by adsorption on activated
montmorillonites
M. Mekhloufi & A. Zehhaf & A. Benyoucef &
C. Quijada & E. Morallon

Received: 4 May 2013 / Accepted: 2 July 2013 / Published online: 8 August 2013
# Springer Science+Business Media Dordrecht 2013

Abstract Sodium montmorillonite (Na-M), acidic Keywords Adsorption isotherms . 8-Quinolinecarboxylic


montmorillonite (H-M), and organo-acidic montmoril- acid . Activated montmorillonite . Kinetics .
lonite (Org-H-M) were applied to remove the herbicide Thermodynamic data
8-quinolinecarboxylic acid (8-QCA). The montmoril-
lonites containing adsorbed 8-QCA were investigated
by transmission electron microscopy, FT-IR spectrosco-
Introduction
py, X-ray diffraction analysis, X-ray fluorescence
thermogravimetric analysis, and physical adsorption of
Water pollution is any human-caused contamination of
gases. Experiments showed that the amount of adsorbed
water that reduces its usefulness to humans and other
8-QCA increased at lower pH, reaching a maximum at
organisms in nature. Pollutants such as pesticides, fertil-
pH 2. The adsorption kinetics was found to follow the
izers, and hazardous chemicals often find their way into
pseudo-second-order kinetic model. The Langmuir
water channels. When water supply is contaminated, it
model provided the best correlation of experimental data
posed a great threat to humans, animals, and plants’ health
for adsorption equilibria. The adsorption of 8-QCA
unless it goes through a costly purification procedure.
decreased in the order Org-H-M > H-M > Na-M. Iso-
Water pollution can be considered in a number of ways,
therms were also used to obtain the thermodynamic
but in simple terms, it is an addition of an array of
parameters. The negative values of ΔG indicated the
unwanted contaminant elements to it (Forster 2003). Pol-
spontaneous nature of the adsorption process.
lution of surface water and groundwater causes a risk to
human health because of the potential health hazards of
their contents of inorganic and organic compounds. Pesti-
M. Mekhloufi : A. Zehhaf : A. Benyoucef (*) cides are a group of hazardous compounds that may
Laboratoire de Chimie Organique, Macromoléculaire et des
pollute water due to their extensive application in agricul-
Matériaux, Université de Mascara,
Bp 763, Mascara 29000, Algeria ture (Ayranci and Hoda 2005; Daneshvar et al. 2007;
e-mail: ghani29000@yahoo.fr Gupta et al. 2006).
Substituted quinolinecarboxylic acids are a class of
C. Quijada
highly selective auxin herbicides that effectively controls
Departamento de Ingeniería Textil y Papelera,
Universitat Politècnica de València, important dicotyledonous weeds, such as cleavers
Pza Ferrandiz i Carbonel, 03801 Alcoy, Alicante, Spain (Galium aparine L.) in sugar beet, oilseed rape, and
wheat (Grossmann and Scheltrup 1998; Sotelo et al.
E. Morallon
2002). Their action mechanism involves the interference
Departamento de Química Física e Instituto Universitario de
Materiales, Universidad de Alicante, with nucleic acid synthesis, which affects the DNA reg-
Apartado 99, 03080 Alicante, Spain ulation during the RNA formation by depression of a
10366 Environ Monit Assess (2013) 185:10365–10375

gene or activation of RNA polymerase, simply affecting montmorillonite, and organo-montmorillonite from aque-
RNA message to proteins. ous solutions. This adsorption study has not been report-
8-Quinolinecarboxylic acid (8-QCA) is a typical per- ed in the literature. We have been interested in the Na
sistent organic pollutant. Because of its persistence, low montmorillonite and acid-activated montmorillonite in-
volatility, bioaccumulation potential, and high toxicity, it tercalated with ammonium acetate via ion exchange. As
is a major environmental issue, drawing much scientific the montmorillonite is progressively intercalated by this
and public attention. The 8-QCA can be transferred modification, the surface properties of the clay mineral
throughout the food chain and pollute the entire ecolog- change considerably, and so does the ability of organo-
ical environment, also posing problems to human health. montmorillonites to remove organic contaminants from
For soils that are contaminated by organic pollutants, water. To understand the adsorption mechanism, the cor-
there are usually two kinds of remediation technologies responding kinetics, isotherm, thermodynamic, pH influ-
available, namely, in situ and ex situ remediations. Ex ence on adsorption, and equilibrium models were stud-
situ remediation may gradually be replaced by in situ ied. UV adsorption spectroscopy was used to analyze the
remediation (Sui et al. 2003). Among the many in situ concentration of 8-QCA before and after adsorption onto
remediation technologies, remediation of contaminated the original and modified clay minerals.
soils and groundwaters by organoclays has been a topic
under development (Hermosín et al. 2006; Nzengung
et al. 1996; Oyanedel-Craver et al. 2006; Özcan et al. Materials and methods
2006; Pernyeszi et al. 2006; Wang et al. 1997; Zhou et al
2006). Organoclays are synthesized by exchanging the Reagents
naturally occurring inorganic cations with organic cat-
ions. By exchanging organic cations on the surface of All reagents used in the present study were of analyt-
clay minerals, a material with radically different physi- ical grade. 8-QCA of 98 % purity was obtained from
cal and chemical properties is produced (Jaynes and Sigma-Aldrich Chemicals. Ammonium acetate of
Boyd 1991; Seki and Yurdakoc 2005). Modification of 98 % purity was obtained from Fluka Ltd. The water
clay minerals by ion exchange with organic cations has employed for the preparation of these pesticide solu-
been shown to have a significant impact on their tions was double-distilled water. The chemical struc-
sorptivity (Klumpp et al. 2004; Richard et al. 2001; ture of 8-QCA used is shown in Fig. 1. All other
Wiles et al. 2005). reactants were also purchased from Sigma-Aldrich
In this study, Na montmorillonite was used as an Chemicals. The raw Na montmorillonite (named as
original clay mineral. Synthesized organoclays are maghnite) used in this study was from Tlemcen (West
organophilic rather than hydrophilic and are able to sorb Algeria).
nonpolar organic contaminants, unlike unmodified clays,
which are poor sorbents for these compounds (Seki and Clay adsorbents
Yurdakoc 2005; Redding et al. 2002; Yasser and Jamal
2004). Because the naturally occurring inorganic cations The clay mineral sample was washed with distilled
in the interlamellar spaces and on the surface of clays are water to remove impurities; the raw montmorillonite
not part of the inter-clay structure, they can easily be (20 g) was crushed for 20 min using the Prolabo
replaced by organic cations (Smith and Galan 1995).
Sorption of hydrophobic organic contaminants by
organoclays can occur via different mechanisms and to
different degrees depending on the type of organic cat-
ions used, the type of original clay and its cation ex-
change capacity (CEC), the proportion of CEC satisfied
by organic cations, and characteristics of sorbate, such as
size, shape, solubility, and hydrophobicity (Redding et al.
2002; Richard et al. 2001; Yasser and Jamal 2004).
The aim of this work is to study the adsorption
of 8-QCA on Na montmorillonite, acid-activated Fig. 1 Structure of herbicide 8-quinolinecarboxylic acid
Environ Monit Assess (2013) 185:10365–10375 10367

ceramic balls grinder. Then, it was dried at 423 K for assess the narrowest micropores (VDR (CO2)) (pore
2 h and stored in tightly stoppered glass bottles for later size smaller than around 0.7 nm) also by application
use (samples of sodium montmorillonite (Na-M)). The of the Dubinin–Radushkevich equation at relative
Na-M sample was activated in acid as acidic montmo- pressures below 0.025 (Cazorla et al. 1996, 1998;
rillonite (H-M) by the procedure of Belbachir and Lozano et al. 2009).
Bensaoula (2001). Briefly, this procedure consists on For transmission electron microscopy (TEM) obser-
refluxing 20 g of Na-M in 200 mL of 0.25 M H2SO4 for vations, the samples were dispersed in water and sup-
3 h. The resulting acidic activated clay was centrifuged ported on TEM grids. The images were collected using
and washed with water several times until it was the JEOL (JEM-2010) microscope, working at an op-
free of SO42−, and the pH of the washing was 6.8. eration voltage of 200 kV.
Finally, the sample was dried at 378 K in air until The X-ray diffraction of the powder nanocompos-
constant weight. This acid activation also removes ites was taken using the Bruker CCD-Apex equipment
sodium from its composition (Salavagione et al. with an X-ray generator (Cu Kα and Ni filter) operated
2008; Zehhaf et al. 2012). The organo-acidic at 40 kV and 40 mA.
montmorillonite (Org-H-M) was prepared by the X-ray fluorescence spectroscopy of the powder
addition of 5 mmol of the solid ammonium acetate samples was made using the Philips PW1480
salt to 1 L of a 1 % (w/v) aqueous dispersion of equipment with the UniQuant II software to deter-
acid-activated montmorillonite under stirring which mine element concentration in a semiquantitative
was continued for 24 h. The Org-H-M was separated by way.
centrifugation (4 g, 30 min). The sediments were The thermogravimetric analysis was carried out
washed three times with distilled water. The chemical using the Mettler Toledo 851E/1600/LG equipment in
composition of the three different clay minerals is in- nitrogen atmosphere at heating rate of 10 °C/min.
cluded in Table 1.
Herbicide (8-QCA) adsorption studies
Characterization of the clay adsorbent
The clay mineral samples were dried at 80 °C
The porous texture of all samples was determined by under vacuum for 24 h before adsorption. Then,
physical adsorption of gases (N2 at 77 K and CO2 at 0.5 g of adsorbent was put in contact with 50 mL
273 K) using an automatic adsorption system of an aqueous solution of 8-QCA with concentra-
(Autosorb-6, Quantrachrome Corporation) after sam- tion ranging from 1 to 1,500 mg L−1 at 25 °C and
ple outgassing at 383 K under vacuum for 4 h. Nitro- for 24 h. Each adsorption experiment was run in
gen adsorption at 77 K was used for determining the duplicate. The pH of the 8-QCA initial solution
total volume of micropores (VDR (N2)) (pore size ranged from 2 to 9.0.
smaller than 2 nm) by applying the Dubinin– The same adsorption experiments were repeated at
Radushkevich (DR) equation (range of relative pres- 298, 308, and 318 K. The equilibrium concentrations
sures used for the DR analysis was 0.005<P/P0<0.17) were measured by using the Unicam UV2-100 Double
and for determining the specific surface area by the Beam Scanning UV–Visible Spectrometer λ=312 nm for
Brunauer–Emmett–Teller (BET) equation (SBET), 8-QCA. The adsorbed amount of herbicide was calcu-
whereas the adsorption of CO2 at 273 K was used to lated by a difference between the initial concentration

Table 1 Composition (in weight percent) of Na-montmorillonite (Na-M), acidic montmorillonite (H-M), and organo-acidic montmo-
rillonite (Org-H-M)

Composition SiO2 Al2O3 MgO Fe2O3 K2O TiO2 Rb2O ZrO2 SO3 ZnO CaO Na2O P2O5 MnO SrO

Na-M 68.01 18.31 3.56 3.19 1.89 0.08 0.02 0.01 0.16 0.02 0.02 4.58 0.02 0.11 0.02
H-M 72.93 20.66 1.46 2.28 2.46 0.18 0.01 – – – – – 0.01 – –
Org-H-M 72.27 20.42 1.56 2.78 2.67 0.27 0.01 – – – – – 0.01 – –
10368 Environ Monit Assess (2013) 185:10365–10375

and that after a time, according to the following surface. The equation is commonly represented
equation: by the following:
1
ðc0 −ct ÞV lnqe ¼ lnK f þ lnce ð4Þ
qt ðmg=gÞ ¼ ð1Þ n
1; 000W
where qe is the amount adsorbed (in milligram per
gram), ce is the equilibrium concentration of the
where qt (in milligram per gram) is the amount of
adsorbate (in milligram per liter), and Kf (in mil-
organic pollutant adsorbed on the adsorbent at time t,
ligram1−1/n per liter1/n per gram) and n are the
c0 is the initial concentration (in milligram per liter), ct
Freundlich constants characteristics of the system,
is the concentration of organic pollutant in solution at
corresponding to the adsorption capacity and the
time t (in milligram per liter), V is the volume of the
strength of adsorption, respectively.
solution used (in milliliter), and W is the weight of the
In order to determine the adsorption kinetics of 8-
adsorbent used (in gram).
QCA herbicide, first- and second-order kinetic models
In order to analyze the experimental data of
were checked. The first-order rate expression (Kul and
organic adsorption from aqueous solutions, the
Koyunchu 2010) is expressed as follows:
linear form of well-known Langmuir isotherm
equation applicable to adsorption up to monolayer dqt
¼ k 1 ðqe −qt Þ ð5Þ
levels on a homogeneous surface with no interac- dt
tion between adsorbed molecules was chosen as
where qe and qt are the amounts of 8-QCA adsorbed
follows:
onto the montmorillonite (in milligram per gram), re-
ce 1 ce spectively, at equilibrium and at time t, and k1 is the
¼ þ ð2Þ first-order rate constant (per minute). After integration
qe K l c m qm
from t=0 to t and from qt=0 to qe, it becomes the
where qe is the amount adsorbed (in milligram per Lagergren rate equation as follows:
gram), ce is the equilibrium concentration of the
k1
adsorbate (in milligram per liter), and qm (in mil- logðqe −qt Þ ¼ logqe − t: ð6Þ
ligram per gram) and Kl (in liter per milligram) are 2:303
Langmuir constants, with qm being the maximum In most cases, the first-order equation of Lagergren
adsorption capacity of adsorbent in a monolayer, does not apply well throughout the whole range of
and Kl is related to the free energy of adsorption contact times and is generally applicable over the ini-
(Gu et al. 1995). tial 20–30 min of the adsorption process.
The essential characteristics of the Langmuir iso- A pseudo-second-order rate law expression was also
therm can be expressed in terms of a dimensionless used; the kinetic rate equation is expressed as (Ho and
constant separation factor Rl as follows: McKay 1999) follows:
1 dqt
Rl ¼ ð3Þ ¼ k 2 ðqe −qt Þ2 ð7Þ
1 þ K l C0 dt
where Co (in milligram per liter) is the highest initial where k2 is the second-order rate constant (in gram per
concentration of adsorbate, and Kl (in liter per milli- milligram per minute). At boundary conditions from t=0 to
gram) is Langmuir constant. t and from qt = 0 to qe, the rate law becomes the following:
The essential characteristics of the Langmuir iso- t 1 1
therm can be expressed in terms of a dimensionless ¼ þ t: ð8Þ
qt k 2 q2e qe
constant separation factor Rl. The value of Rl indicates
the shape of the isotherm to be either unfavorable
(Rl >1), linear (Rl =1), favorable (0 < Rl < 1), or irre- Adsorption thermodynamics
versible (Rl = 0) (Salman and Hameed 2010).
The Freundlich isotherm is an empirical equa- The thermodynamic parameters of the adsorption, i.e.,
tion based on adsorption on a heterogeneous the standard enthalpy ΔH°, Gibbs free energy ΔG°,
Environ Monit Assess (2013) 185:10365–10375 10369

and entropy ΔS° were calculated using the following structure. In the case of Na-M clay, it is possible to
equations: determine the separation between layers being around
3.6 Å. However, clear differences can be observed differ-
ΔG ¼ −RT lnK l ð9Þ
ences in the morphology of both samples with respect to
that observed for Org-H-M clay (Fig. 2c). This latter clay
ΔS  ΔH 
lnK l ¼ − ð10Þ mineral shows a largely exfoliated morphology in agree-
R RT
ment with the nitrogen isotherms. The specific surface
ΔG ¼ ΔH  −T :ΔS  ð11Þ area of the samples was obtained from the BET method
by using the adsorption data. Table 2 includes the porous
where R is the ideal gas constant (in joule per mole per texture characterization data of the three samples. It can be
Kelvin), Kl is the equilibrium constant, and T is observed that the specific surface area increases from
the temperature (in Kelvin); ΔH° and ΔS° values 32 m2 g−1 for the Na-M sample to 140 m2 g−1 for the
can be obtained from the slope and intercept, H-M and to 161 m2 g−1 for Org-H-M sample. The micro-
respectively, of van’t Hoff plots of Kl obtained pore and mesopore volumes also increase in the Org-H-M
from the Langmuir isotherm versus 1/T. > H-M > Na-M samples, possibly as a consequence of the
dissolution of the exchangeable cations like sodium and
the partial dissolution of the structural cations like Fe3+
Results and discussion and Mg2+ (Table 1) (Noyan et al. 2007).
The TG experiments (not shown) contain the
Characterization of the adsorbents typical features for montmorillonites, namely, two
main processes at around 85 and 450 °C. The
The nitrogen adsorption/desorption isotherms at 77 K for first one corresponds to the evolution of weakly
natural and acidic activated samples (not shown) exhibit a bonded water molecules with a loss between 0.5
shape similar to the type II isotherms, with a hysteresis and 2 % by weight, due to the loss of interlayer
loop of the H3 type according to IUPAC (Sing et al. water. The second one can be associated with
1985). This shape indicates that the samples are mainly dehydroxylation of the octahedral sheet (Huang
mesoporous solids but also contain some micropores. The et al. 2004). Above 650 °C, the clay minerals
hysteresis loop appears in porous materials with slit- are completely calcined.
shaped pores or pores with narrow necks and wide bodies,
and then, they can be associated with capillary condensa- Herbicide (8-QCA) adsorption
tion of liquid nitrogen in mesopores. The hysteresis phe-
nomenon becomes more prominent after acidic and or- Figure 3 shows the adsorption isotherms of herbicide
ganic activation because the amount of mesopores (8-QCA) on the different bentonites in solutions of
increases by the structural deformation and reflects a initial pH=2. The modified clay minerals, Org-H-M
change from the laminar structure of the Na montmoril- and H-M, displayed a higher adsorbed amount than
lonite to a delaminated structure (Temuulin et al. 2004). Na-M. The adsorption isotherms on both Org-H-M and
Figure 2 contains the TEM images of the Na-M (Fig. 2a) H-M had an L-type profile. This profile of isotherms
and H-M (Fig. 2b) samples which show a laminar suggests that the adsorption is driven by the affinity

Fig. 2 TEM images of a Na-montmorillonite, b acidic montmorillonite, and c organo-acidic montmorillonite


10370 Environ Monit Assess (2013) 185:10365–10375

Table 2 Textural characterization of Na montmorillonite adsorbed amounts and correspond to the adsorption
(Na-M), acidic montmorillonite (H-M), and organo-acidic
isotherms plateaus. The 8-QCA isotherms exhibited
montmorillonite (Org-H-M)
classical type I characteristics and, as such, obeyed
Samples SBET VDR (N2) VDR (CO2) Vmeso the Langmuir equation. The high levels of adsorption
m2 g−1 cm3 g−1 cm3 g−1 cm3 g−1 indicate that the 8-QCA is attracted sufficiently to the
Na-M 32 0.09 0.01 0.16
hydrophobic surfaces of organo-montmorillonites, and
eventually, it must have penetrated into interlayer
H-M 140 0.19 0.02 0.69
spaces (Yapar et al. 2005).
Org-H-M 161 0.22 0.03 0.70
There are two types of interaction between polar organ-
ic compounds and organoclays: adsorption and partition.
between the 8-QCA molecules and the hydrophobic Surface adsorption may include ion exchange, protonation,
interlayer space (Celis et al. 2007) due to the interca- hydrogen bonding, and coordination/ion–dipole reactions
lation of ammonium acetate and acid-activated mont- with clays (Bleam 1990; Sing et al. 1985). Partition in-
morillonite. In comparison, the adsorption by Na-M, volves interactions between organic matter and nonionic
although noticeable, was lower than that by the mod- organic compounds. When a large organic molecule, for
ified montmorillonites. The maximum adsorption ca- example, ammonium acetate, interacts with clay minerals,
pacity decreases in the order Org-H-M > H-M > Na-M. the partition process plays an important role. For this
The Org-H-M shows the highest adsorption capacity in mechanism, the adsorption of organic contaminants is
agreement with the increase in the BET surface area functionally and conceptually similar to the dissolution of
values. organic contaminants in a bulk-phase organic solvent. This
The Langmuir and Freundlich isotherms were used mechanism results from the change of surface properties
to describe the experimental results. From the linear from hydrophilic to hydrophobic because the long tails of
plot of specific adsorption (Ce/qe) against the equilib- organic molecules form a hydrophobic phase in the
rium concentration (Ce) (figures not shown), the Lang- interlayer of the organoclays (Boyd et al. 1988; Smith
muir constants qm and Kl were determined from the et al. 1990).
slope and intercept and are listed in Table 3. It showed The dimensionless separation factors calculated for
that the model which gives a good fit to the experi- 8-QCA adsorption at 20 °C (Eq. 3) are as follows:
mental data is the Langmuir model. The values of Rl=0.19 for adsorption of 8-QCA on Na-M, Rl=0.16
maximum adsorption capacity determined using Lang- for adsorption of 8-QCA on H-M, and Rl=0.08 for
muir expression are higher than the experimental adsorption of 8-QCA on Org-H-M. Rl values were less

Fig. 3 The isotherms of 8-


QCA adsorption from aque-
ous solutions on the mont-
morillonites (Na-M, H-M,
and Org-H-M) after equilib-
rium time 24 h at 25 °C
Environ Monit Assess (2013) 185:10365–10375 10371

Table 3 Freundlich and Lang-


muir coefficients obtained from Isotherms Adsorbents Coefficients
the adsorption isotherms of 8-
QCA on the montmorillonite Freundlich Kf (mg1−1/nL1/ng−1) n R2
samples at 298 K and pH 2 Na-M 2.40 2.09 0.87
H-M 3.87 2.33 0.96
Org-H-M 8.79 2.98 0.93
Langmuir qm (mg g−1) Kl (L mg−1)×10−3 R2
Na-M 65.4 0.47 0.98
H-M 67.6 0.73 0.98
Org-H-M 75.9 1.71 0.99

than 1 and greater than 0, indicating favorable adsorp-


tion for the three processes.
Figure 4 shows XRD diffraction patterns of all
samples used (before and after adsorption of herbi-
cide). The intercalation of Na montmorillonite by pro-
ton cations (H-M) and, subsequently, by ammonium
acetate (Org-H-M) changes the basal spacing from
d(001)=13.10 to 15.58Ǻ and to 13.02Ǻ, respectively.
A new change to basal spacing for Na-M/8-QCA, H-
M/8-QCA, and Org-H-M/8-QCA was observed after
adsorption of herbicide (Table 4). These results con-
firm that all interlayer spaces were expanded by herbi-
cide 8-QCA.
The infrared spectra of pure 8-QCA, H-M, Org-H-M,
and the herbicide 8-QCA intercalated in Org-H-M are
shown in Fig. 5. The common features to all the clay
minerals are the band at 3,625 cm−1, the broad band at
3,380 cm−1, the intense band at 1,008 cm−1, and the
sharp bands at 798 and 515 cm−1. All these features
correspond to characteristic vibrational modes of the
octahedral and tetrahedral sheets in a montmorillonite

Table 4 Peak maximum and d spacing of modified montmoril-


lonites before and after adsorption of 8-QCA herbicide

Samples Peak maximum, Basal spacing,


2θmax (°) d(001) (Å)

Before
Na-M 6.65 13.10
H-M 5.59 15.58
Org-H-M 6.69 13.02
After
Na-M/8-QCA 5.66 15.39
H-M/8-QCA 5.79 15.05
Fig. 4 XRD diffraction patterns of montmorillonites Na-M, H-M, Org-H-M/8-QCA 5.65 15.42
and Org-H-M before and after adsorption of 8-QCA herbicide
10372 Environ Monit Assess (2013) 185:10365–10375

and ν(Si–O–(Mg)Al) modes (515 cm−1) (Dixon et al.


1977; do Nascimento et al. 2004; Lee et al. 2003; Wu
et al. 2008). The intensity of the bands at 3,380 and
1,631 cm−1 indicated a relatively high water content,
suggesting that the replacement of the original interlayer
inorganic cations of the clay and their hydration occurred
to a great extent. The spectrum of Org-H-M shows an
additional absorption the two N–H-bound vibrations at
1,443 cm−1 (N–H bending) and from 2,920 to 2,840 cm−1
(N–H stretching). The spectrum of Org-H-M/8-QCA
shows the increase in band intensity in the region between
1,615 and 1,650 cm−1 assigned to stretching of C=N
bonds and adsorbed water molecules and the band at
2,920 cm−1 related to N–H stretching. One singular band
around 1,548 cm−1 can be observed in spectra of mont-
morillonite with herbicide. These bands corroborate the
protonation of the nitrogen atoms of the herbicide,
confirming the interaction between herbicide 8-QCA
and montmorillonites.

Effect of the solution pH

To study the influence of the pH on the adsorption


capacity of Na-M, H-M, and Org-H-M, experiments
Fig. 5 FT-IR spectra of H-M, Org-H-M, Org-H-M/8-QCA, and
pure 8-QCA herbicide were run at various pHs ranging from 1 to 9. At first,
the uptake of 8-QCA increased slowly with the increas-
ing pH, reaching a maximum for pH 2 and decreased
layer: ν(OH) from structural OH groups in again at higher pHs (Fig. 6). It was observed that the
dioctahedral smectite sheets (3,625 cm−1), ν(Si–O) adsorption is highly dependent on the pH of the solu-
modes (1,008 cm−1), ν(Al(Mg–O–)) mode (798 cm−1), tion, which affects the surface charge of the adsorbent

Fig. 6 Effect of pH on 8-
QCA adsorption on Na-M,
H-M, and Org-H-M. Initial
concentration of 8-
QCA=1,500 mg L−1, ad-
sorption time=24 h
Environ Monit Assess (2013) 185:10365–10375 10373

Table 5 Comparison of the first- and second-order adsorption rate constants, for calculated (qe, cal) and experimental (qe, exp) values at 8-
QCA of concentration 1,500 mg L−1 and pH 2

Adsorbent qe, exp First-order kinetic model Second-order kinetic model

mg g−1 k1 qe, cal R2 k2, ads qe, cal R2


min−1 mg g−1 g mg−1 min−1 mg g−1

Na-M 54.7 0.019 85.9 0.90 0.0089 60.9 0.97


H-M 59.5 0.019 91.2 0.91 0.0085 66.2 0.97
Org-H-M 69.9 0.014 87.9 0.95 0.0071 77.5 0.97

and the degree of ionization and speciation of the with the correlation factor R2 for different 8-QCA concen-
adsorbate. trations, obtained after applying the above mentioned
This was due to the chemical characteristics of the Na- models to the experimental data. It can be observed that
M, H-M, and Org-H-M with acidic pHpzc of 2.9, 3.7, and the second-order model fits better the experimental data in
4.5, respectively. At a solution pH lower than the pHpzc, the whole range of 8-QCA concentrations used (from 1 to
the total surface charge would be on average positive, 1,500 mg L−1).
whereas at a higher solution pH, it would be negative. The second-order model is based on the assumption
The 8-QCA uptake was the highest at lower pH where that the rate-limiting step may be a chemical adsorption
the pH was below the pKa of 8-QCA (8-QCA shows two involving valence forces through sharing or exchange
pKa values, 2.4 and 6.8). The first one, corresponding to of electrons between adsorbent and adsorbate (Brigatti
the deprotonation of the carboxylic group, is comparable et al. 2000).
with that observed in picolinic acid. The value of 6.8 is
due to the dissociation of the NH+ quinoline group Thermodynamic studies
(Garribba et al. 2003; Kiss et al. 2000). At acidic pH,
the herbicide was undissociated, and the dispersion inter- The effect of temperature on adsorption capacity of Na-
actions predominated (Hamdaoui and Naffrechoux M, H-M, and Org-H-M was studied by carrying out a
2007). At basic pH, however, the lower adsorption is series of isotherms at 298, 308, and 318 K using an
due to the repulsion between the negatively charged initial 8-QCA concentration 1,500 mg L−1 at pH 2. The
surface of montmorillonite samples and the anions of effect of temperature on the equilibrium adsorption ca-
8-QCA. pacity of clay mineral samples is presented in Table 6.
The negative value for the Gibbs free energy for 8-QCA
Adsorption kinetics adsorption on the three clay minerals shows that the
adsorption process is spontaneous and that the degree
The kinetics of the herbicide 8-QCA batch adsorption on of spontaneity of the reaction increases with increasing
the three samples used in this work was examined in terms temperature. The overall adsorption process seems to be
of pseudo-first-order and pseudo-second-order kinetics. endothermic (ΔH°Na-M=19.88 kJ mol−1, ΔH°H-
−1 −1
Table 5 includes the relevant kinetic parameters, together M=14.56 kJ mol , and ΔH°Org-H-M=15.73 kJ mol ).

Table 6 Thermodynamic constants for the adsorption of 8-QCA on Na-M, H-M and Org-H-M montmorillonites at various tempera-
tures, (initial 8-QCA concentration 1,500 mg L−1 and pH 2)

Adsorbents Na-M H-M Org-H-M

T (K) 298 308 318 298 308 318 298 308 318
−1
ΔG (kJ mol ) −4.34 −5.15 −5.96 −3.81 −4.43 −5.05 −2.89 −3.52 −4.14
ΔH (kJ mol−1) 19.88 14.56 15.73
ΔS (J mol−1 K−1) 81.27 61.65 62.49
10374 Environ Monit Assess (2013) 185:10365–10375

Although not very high, these values of ΔH° can be Cazorla, A. D., Alcañiz, M. J., & Linares, S. A. (1996). Charac-
terization of activated carbon fibers by CO2 adsorption.
interpreted on the basis of a considerably strong inter-
Langmuir, 12, 2820–2824.
action between 8-QCA and montmorillonite surface. Cazorla, A. D., Alcañiz, M. J., De la Casa, L. M. A., & Linares,
Table 6 also shows that the ΔS° value was positive. S. A. (1998). CO2 as an adsorptive to characterize carbon
This occurs as a result of redistribution of energy be- molecular sieves and activated carbons. Langmuir, 14,
4589–4596.
tween the adsorbate and the adsorbent. Adsorption is
Celis, R., Trigo, C., Facenda, G., Hermosín, M. C., & Cornejo, J.
thus likely to occur spontaneously at normal and high (2007). Selective modification of clay minerals for the
temperatures because ΔH>0 and ΔS>0. adsorption of herbicides widely used in olive groves. Jour-
nal of Agricultural and Food Chemistry, 55, 6650–6658.
Daneshvar, N., Aber, S., Khani, A., & Rasoulifard, M. H. (2007).
Investigation of adsorption kinetics and isotherms of
Conclusion imidacloprid as a pollutant from aqueous solution by ad-
sorption onto industrial granular activated carbon. Journal
of Food Agriculture and Environment, 5, 425–429.
It has been demonstrated here that Na montmorillonite Dixon, J. B., Weed, S. B., & Dinauer, R. C. (1977). Minerals in
modified by the addition of the ammonium acetate soil environments. Berkeley: Soil Science Society of
exhibits interesting adsorption properties with regard America.
to 8-QCA herbicide. The organic compound surfactant Do-Nascimento, G. M., Constantino, V. R. L., Landers, R., &
Temperini, M. L. A. (2004). Aniline polymerization into mont-
has a dual effect. First, it increases the affinity of the morillonite clay: a spectroscopic investigation of the interca-
microporous interlayer adsorption sites. In the case of lated conducting polymer. Macromolecules, 37, 9373–9385.
8-QCA adsorption, this improved affinity allows ad- Forster, C. F. (2003). Wastewater treatment and technology.
sorption to occur much more efficiently than in the case London: Thomas Telford.
Garribba, E., Micera, G., Sanna, D., & Chruscinska, E. L.
of conventional Na montmorillonite and acid-activated (2003). Oxovanadium(IV) complexes of quinoline deriva-
montmorillonite. Second, surfactant molecules create tives. Inorganica Chimica Acta, 348, 97–106.
new adsorption sites on the external surfaces of the Grossmann, K., & Scheltrup, F. (1998). Studies on the mecha-
clay layers. These surface chemical effects may be very nism of selectivity of the auxin herbicide quinmerac. Pesti-
cide Science, 52, 111–118.
useful in designing a viable adsorbent for future Gu, B., Schmitt, J., Chen, Z., Liang, L., & McCarthy, J. F.
groundwater treatment applications. (1995). Adsorption and desorption of different organic mat-
ter fractions on iron oxide. Geochimica et Cosmochimica
Acta, 59, 219–229.
Acknowledgments This work was supported by the National
Gupta, V. K., Ali, I., & Saini, V. K. (2006). Adsorption of 2,4-D
Agency for the Development of University Research (CRSTRA)
and carbofuran pesticides using fertilizer and steel industry
and the Directorate General of Scientific Research and Techno-
wastes. Journal of Colloid and Interface Science, 299, 556–
logical Development of Algeria. The Ministerio de Economía y
563.
Competitividad (MAT2010-15273 project) and FEDER are also
Hamdaoui, O., & Naffrechoux, E. (2007). Modeling of adsorp-
acknowledged.
tion isotherms of phenol and chlorophenols onto granular
activated carbon. Part II. Models with more than two pa-
rameters. Journal of Hazardous Materials, 147, 401–411.
Hermosín, M. C., Celis, R., Facenda, G., Carrizosa, M. J.,
References Ortega-Calvo, J. J., & Cornejo, J. (2006). Bioavailability
of the herbicide 2,4-D formulated with organoclays. Soil
Biology and Biochemistry, 38, 2117–2124.
Ayranci, E., & Hoda, N. (2005). Adsorption kinetics and iso- Ho, Y. S., & McKay, G. (1999). Pseudo-second-order model for
therms of pesticides onto activated carbon-cloth. sorption processes. Process Biochemistry, 34, 451–465.
Chemosphere, 60, 1600–1607. Huang, F. C., Lee, F. J., Lee, C. K., & Chao, H. P. (2004). Effects
Belbachir, M., Bensaoula, A. (2001). US Patent no. 6, 274, 527 of cation exchange on the pore and surface structure and
B1. adsorption characteristics of montmorillonite. Colloid Sur-
Bleam, W. F. (1990). The nature of cation-substitution sites in face A, 239, 41–47.
phyllosilicates. Clays and Clay Minerals, 38, 527–536. Jaynes, W. F., & Boyd, S. A. (1991). Clay mineral type and organic
Boyd, S. A., Mortland, M. M., & Chiou, C. T. (1988). compound sorption by hexadecyltrimethylammonium-
Sorption characteristics of organic compounds on exchanged clays. Soil Science Society of American Journal,
hexadecyltrimethylammonium-smectite. J Soil Sci Soc 55, 43–48.
Am, 52, 652–657. Kiss, E., Petrohan, K., Sanna, D., Garribba, E., Micera, G., &
Brigatti, M. F., Lugli, C., & Poppi, L. (2000). Kinetics of Kiss, T. (2000). Solution speciation and spectral studies on
heavy-metal removal and recovery in sepiolite. Applied oxovanadium(IV) complexes of pyridinecarboxylic acids.
Clay Science, 16, 45–57. Polyhedron, 19, 55–61.
Environ Monit Assess (2013) 185:10365–10375 10375

Klumpp, E., Ortega, C. C., & Klahre, P. (2004). Sorption of 2,4- Sing, K., Everet, D., Haul, R., Moscou, L., Pierotty, R.,
dichlorophenol on modified hydrotalcites. Colloids and Rouquerol, J., et al. (1985). Reporting physisorption data
Surfaces A: Physicochemical and Engineering Aspects, for gas/solid systems with special reference to the determi-
230, 111–116. nation of surface area and porosity. Pure and Applied
Kul, A. R., & Koyunchu, H. (2010). Heavy metal removal from Chemistry, 57, 603–619.
municipal solid waste fly ash by chlorination and thermal Smith, J., Jaffe, P., & Chiou, C. (1990). Effect of ten quaternary
treatment. Journal of Hazardous Materials, 179, 332–339. ammonium cations on tetrachloromethane sorption to clay
Lee, D. K., Char, K., Lee, S. W., & Park, Y. W. (2003). Structural from water. Environmental Science and Technology, 24,
changes of polyaniline/montmorillonite nanocomposites 1167–1172.
and their effects on physical properties. Journal of Mate- Smith, J. A., & Galan, A. (1995). Sorption of nonionic organic
rials Chemistry, 13, 2942–2947. contaminants to single and dual organic cation bentonites
Lozano, C. D., Suárez, G. F., Cazorla, A. D., & Linares, S. A. from water. Environmental Science and Technology, 29,
(2009). Porous texture of carbons. In F. Beguin & E. 685–692.
Frackowiak (Eds.), Carbons for electrochemical energy stor- Sotelo, J. L., Ovejero, G., Delgado, J. A., & Martínez, I. (2002).
age and conversion systems (pp. 115–162). Boca Raton: CRC. Comparison of adsorption equilibrium and kinetics of four
Noyan, H., Onal, M., & Sarikaya, Y. (2007). The effect of chlorinated organics from water onto GAC. Water
sulphuric acid activation on the crystallinity, surface area, Research, 36, 599–608.
porosity, surface acidity, and bleaching power of a bentonite. Sui, H., Li, X. G., Huang, G. Q., Zhang, Y., & Gao, X. F. (2003).
Food Chemistry, 105, 156–163. The in-situ remediation technologies for soils contaminated
Nzengung, V. A., Voudrias, E. A., Nkedi-Kizza, P., Wampler, J. M., by organic chemicals. Techniques and Equipment for Envi-
& Weaver, C. E. (1996). Organic cosolvent effects on sorption ronmental Pollution Control, 4, 41–45.
equilibrium of hydrophobic organic chemicals by organoclays. Temuulin, J., Jadambaa, T. S., Burmaa, G., Erdenechimeg, S. H.,
Environmental Science and Technology, 30, 89–96. Amarsanaa, J., & MacKenzie, K. J. D. (2004). Characteri-
Oyanedel-Craver, V. A., Fuller, M., & Smith, J. A. (2006). sation of acid activated montmorillonite clay from Tuulant
Simultaneous sorption of benzene and heavy metals onto (Mongolia). Ceramics International, 30, 251–255.
two organoclays. Journal of Colloid and Interface Science, Wang, X. R., Wu, S. N., & Li, W. S. (1997). Contaminated
309, 485–492. environment remediation with organoclay minerals. Envi-
Özcan, A., Ömeroglu, C., Erdogan, Y., & Özcan, A. S. (2006). ronmental Chemistry, 16, 1–14.
Modification of bentonite with a cationic surfactant: an Wiles, M. C., Huebner, H. J., McDonald, T. J., Donnelly, K. C.,
adsorption study of textile dye reactive blue 19. Journal of & Phillips, T. D. (2005). Matrix-immobilized organoclay
Hazardous Materials, 140, 173–179. for the sorption of polycyclic aromatic hydrocarbons and
Pernyeszi, T., Kasteel, R., Witthuhn, B., Klahre, P., Vereecken, pentachlorophenol from groundwater. Chemosphere, 59,
H., & Klumpp, E. (2006). Organoclays for soil remediation: 1455–1464.
adsorption of 2,4-dichlorophenol on organoclay/aquifer Wu, C. S., Huang, Y. J., Hsieh, T. H., Huang, P. T., Hsieh, B. H.,
material mixtures studied under static and flow conditions. Han, Y. K., et al. (2008). Studies on the conducting nano-
Applied Clay Science, 32, 179–189. composite prepared by in situ polymerization of aniline
Redding, A. Z., Burns, S. E., Upson, R. T., & Anderson, E. F. monomers in a neat (aqueous) synthetic mica clay. Journal
(2002). Organoclay sorption of benzene as a function of of Polymer Science Part A: Polymer Chemistry, 46, 1800–
total organic carbon content. Journal of Colloid and Inter- 1809.
face Science, 250, 261–264. Yapar, S., Ozbudak, V., Dias, A., & Lopes, A. (2005). Effect of
Richard, W. G., Walter, J., & Weber, J. R. (2001). Evaluation of adsorbent concentration to the adsorption of phenol on
shale and organoclays as sorbent additives for low- hexadecyltrimethylammonium-bentonite. Journal of Haz-
permeability soil containment barriers. Environmental Sci- ardous Materials B, 121, 135–139.
ence and Technology, 35, 1523–1530. Yasser, Z. N., & Jamal, M. S. (2004). Adsorption of phenanthrene
Salavagione, H. J., Amorós, D. C., Tidjane, S., Belbachir, M., on organoclays from distilled and saline water. Journal of
Benyoucef, A., & Morallon, E. (2008). Effect of the intercalated Colloid and Interface Science, 269, 265–273.
cation on the properties of poly(o-methylaniline)/maghnite clay Zehhaf, A., Benyoucef, A., Berenguer, R., Quijada, C., Taleb, S.,
nanocomposites. European Polymer Journal, 44, 1275–1284. & Morallon, E. (2012). Lead ion adsorption from aqueous
Salman, J. M., & Hameed, B. H. (2010). Adsorption of 2,4- solutions in modified Algerian montmorillonites. Journal
dichlorophenoxyacetic acid and carbofuran pesticides onto of Thermal Analysis and Calorimetry, 110, 1069–1077.
granular activated carbon. Desalination, 256, 129–135. Zhou, Q., Frost, R. L., He, H. P., & Xi, Y. F. (2006). Changes in
Seki, Y., & Yurdakoc, K. (2005). Paraquat adsorption onto clays the surfaces of adsorbed para-nitrophenol on HDTMA
and organoclays from aqueous solution. Journal of Colloid organoclay. The XRD and TG study. Journal of Colloid
and Interface Science, 287, 1–5. and Interface Science, 307, 50–55.

You might also like