You are on page 1of 14

Journal of Colloid and Interface Science 310 (2007) 411–424

www.elsevier.com/locate/jcis

Adsorptive accumulation of Cd(II), Co(II), Cu(II), Pb(II), and Ni(II)


from water on montmorillonite: Influence of acid activation
Krishna G. Bhattacharyya ∗ , Susmita Sen Gupta
Department of Chemistry, Gauhati University, Guwahati, Assam, India
Received 18 November 2006; accepted 26 January 2007
Available online 26 March 2007

Abstract
The present work investigates the influence of acid activation of montmorillonite on adsorption of Cd(II), Co(II), Cu(II), Ni(II), and Pb(II) from
aqueous medium and comparison of the adsorption capacities with those on parent montmorillonite. The clay–metal interactions were studied
under different conditions of pH, concentration of metal ions, amount of clay, interaction time, and temperature. The interactions were dependent
on pH and the uptake was controlled by the amount of clay and the initial concentration of the metal ions. The adsorption capacity of acid-activated
montmorillonite increases for all the metal ions. The interactions were adsorptive in nature and relatively fast and the rate processes more akin
to the second-order kinetics. The adsorption data fitted both Langmuir and Freundlich isotherms, indicating that strong forces were responsible
for the interactions at energetically nonuniform sites. The Langmuir monolayer capacity of the acid-activated montmorillonite is more than that
of the parent montmorillonite (Cd(II): 32.7 and 33.2 mg/g; Co(II): 28.6 and 29.7 mg/g; Cu(II): 31.8 and 32.3 mg/g; Pb(II): 33.0 and 34.0 mg/g;
and Ni(II): 28.4 and 29.5 mg/g for montmorillonite and acid-activated montmorillonite, respectively). The thermodynamics of the rate processes
showed the adsorption of Co(II), Pb(II), and Ni(II) to be exothermic, accompanied by decreases in entropy and Gibbs free energy, while the
adsorption of Cd(II) and Cu(II) was endothermic, with an increase in entropy and an appreciable decrease in Gibbs free energy. The results have
established the potential use for montmorillonite and its acid-activated form as adsorbents for Cd(II), Co(II), Cu(II), Ni(II), and Pb(II) ions from
aqueous media.
© 2007 Elsevier Inc. All rights reserved.

Keywords: Montmorillonite; Acid-activated montmorillonite; Adsorption kinetics; Adsorption isotherm; Temperature; Enthalpy; Metal ions

1. Introduction toms of Cu(II) poisoning, particularly after drinking beverages


stored in copper or untinned brass containers [3]; low levels of
The multifold increase in the use of heavy metals over the Pb(II) have been identified with anemia, as it causes injury to
past few decades has inevitably resulted in an increased flux the blood-forming system, while high levels cause severe dys-
of metallic substances into the environment. Under favorable function of the kidneys, the liver, and the central and peripheral
pH and other conditions, metals go into solution in water, con- nervous system and high blood pressure [4]; the respiratory
taminating natural water bodies. Cd(II), Co(II), Cu(II), Pb(II), tract and immune systems are sensitive targets of Ni(II) toxi-
and Ni(II) are among the toxic metals: Cd(II) is known to cause city, as it causes chronic bronchitis, emphysema, and impaired
lung insufficiency, bone lesions, and hypertension [1]; Co(II) lung function, and the immune response to nickel is elicited as
affects the respiratory system, the kidneys, and the gastroin- allergic contact dermatitis in humans [5]. Most of these metals
testinal tract, and Co dermatitis and sensitization as a result have been known as carcinogens or co-carcinogens.
of dermal exposure have been well documented [2]; gastroin- Considering the threat of these toxic metals to humans and
testinal distress is one of the reported health effects of Cu(II), all other living beings, an appropriate solution must be found
and vomiting, nausea and abdominal pain are common symp- either by removing them completely from water or by bring-
ing them below the maximum permissible level. A good num-
* Corresponding author. Fax: +91 361 2700311. ber of technologies are available for removal of heavy metals
E-mail address: krishna2604@sify.com (K.G. Bhattacharyya). from solution, namely reverse osmosis, chemical precipitation,
0021-9797/$ – see front matter © 2007 Elsevier Inc. All rights reserved.
doi:10.1016/j.jcis.2007.01.080
412 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

ion exchange, solvent extraction, etc. But adsorption has be- Cr(VI) from water [29]. The adsorption was strongly dependent
come a preferred method for removal, recovery, and recycling on the pH of the medium, with Cr(VI) uptake increasing from
of toxic heavy metals from wastewater [6]. Different conven- pH 1.0 to 7.0, after which the uptake decreases and the process
tional and nonconventional adsorbents have been reported for attains equilibrium within 240 min. The Langmuir monolayer
removal of various metal ions, viz., red mud [7], activated car- capacity of the clay adsorbents was in the range from 10.6 to
bon [8], tree fern [9], sewage sludge [10], sawdust [11], sil- 13.9 mg/g.
ica [12], bone char [13], rice husk [14], bagasse fly ash [15], The present work has been undertaken to investigate the use
resin [16], polymetallic sea nodules [17], modified zeolite [18], of montmorillonite and its acid-activated form for removal of
spirogyra bioadsorbent [19], etc. Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) from aqueous solution
The clay minerals in soils play the role of natural scavengers by adsorption under various environmental conditions in single-
in removing and accumulating contaminants in water passing batch systems.
through the soil by ion exchange and adsorption. High spe-
cific surface area, chemical and mechanical stability, layered 2. Materials and method
structure, high cation exchange capacity (CEC), tendency to
hold water in the interlayer sites, and the presence of Brøn- 2.1. Clay adsorbent
sted and Lewis acidity have made clays excellent adsorbent
materials [20,21]. A good number of studies describing the ad- The parent clay, montmorillonite, SWy-2 (M1), was ob-
sorption of various metals onto natural and modified clays have tained from the University of Missouri-Columbia, Source Clay
been reported. Mellah and Chegrouche [22] used natural ben- Minerals Repository, USA. The clay was modified by treating
tonite for uptake of zinc from aqueous solution and reported it with 0.25 M H2 SO4 (E. Merck, Mumbai, India) to acid-
the maximum Langmuir monolayer capacity as 52.91 mg/g activated montmorillonite (M2) using the procedure of Espan-
for an initial Zn(II) concentration of 300 mg/L at 293 K. taleon et al. [30]. A quantity of 20 g of montmorillonite was
The bentonite–Zn(II) interactions were exothermic and were refluxed with 200 ml of 0.25 M H2 SO4 for 3 h. The resulting
favored by high initial Zn(II) concentration and decreased par- activated clay was centrifuged and washed with water several
ticle size of clay adsorbent. The removal of Cr(III), Cr(VI), and times till it was free of SO2−
4 and dried at 383 K in an air oven
Ag(I) by bentonite was reported by Khan et al. [23], where until constant weight was attained. Both the clays were calcined
the adsorption of Cr(III) and Ag(I) onto bentonite was found at 773 K for 10 h before they were used as adsorbents.
to be exothermic and Cr(VI)–bentonite interactions were en-
dothermic. Sepiolite has been used as an adsorbent for Cd(II), 2.2. XRD measurement
Cu(II), and Zn(II) [24]. Yuvaz et al. [25] used raw kaolin for
removing Mn(II), Co(II), Ni(II), and Cu(II) from aqueous so- A Phillips Analytical X-ray spectrometer (PW 1710) using
lution, where the adsorption affinity was in the order Cu(II) > CuKα radiation was used to characterize the adsorbents.
Ni(II) > Co(II) > Mn(II) and the monolayer adsorption capac-
ity was 10.78, 1.67, 0.92, and 0.45 mg/g, respectively, at 298 K. 2.3. FTIR measurement
Alvarez-Ayuso and Garcia-Sanchez [26] used natural and Na-
exchanged bentonites for the removal of Cr(III), Ni(II), Zn(II), FTIR spectra of the adsorbents were taken with a Perkin–
Cu(II), and Cd(II) from aqueous solution. The authors reported Elmer FTIR Spectrometer (Model Spectrum RXI, range 4400–
that the Langmuir model was most suitable for explaining the 440 cm−1 ).
results and the maximum adsorption capacity was 49.8, 24.2,
23.1, 30.0, and 26.2 mg/g, respectively, for Cr(III), Ni(II), 2.4. Surface area
Zn(II), Cu(II), and Cd(II). The process, favored by increasing
pH, was shown to be very rapid, with >95% uptake in 30 min. The surface areas of the clay adsorbents were estimated fol-
Strawn et al. [27] reported the adsorption of Cu(II) by smectites lowing Sears’ method [31]. A quantity of 0.5 g of an adsorbent
(montmorillonite and bedellite) as a function of varying ionic was acidified with 0.1 N HCl to pH 3–3.5. The volume was
strength and pH. The X-ray adsorption fine structure (XAFS) made up to 50 ml with distilled water after addition of 10.0 g of
and electron paramagnetic resonance (EPR) spectroscopic ex- NaCl. The titration was carried out with standard 0.1 M NaOH
periments on Cu(II) adsorbed on clays revealed that Cu–Cu in a thermostatic bath at 298 ± 0.5 K to pH 4.0, and then to
linkages in the multinuclear complexes were 2.65 Å apart, hav- pH 9.0. The volume, V , required to raise the pH from 4.0 to 9.0
ing a coordination number near one. Lin and Juang [28] tried was noted and the surface area was computed from the follow-
to modify the adsorptive properties of montmorillonite by in- ing equation:
troducing sodium dodecyl sulfate before Cu(II) and Zn(II) ad-
sorption. The modified clay had a higher affinity toward the S (m2 /g) = 32V − 25. (1)
metal ions and equilibrium was attained within 120 min. The
interactions were endothermic with pseudo-first-order kinetics, 2.5. Cation exchange capacity
yielding rate constants of 6.64 × 10−4 and 3.14 × 10−3 min−1
at 298 K for Zn(II) and Cu(II), respectively. Kaolinite and its The copper bis-ethylenediamine complex method has been
modified forms have been successfully used for adsorption of used to estimate the CEC of the clays [32]. A quantity of 50 ml
K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424 413

of 1 M CuCl2 solution was mixed with 102 ml of 1 M ethyl- 2.7.2. Kinetics


enediamine solution to allow the formation of the [Cu(en)2 ]2+ Clay 2 g/L, metal ion concentration 50 mg/L, temperature
complex. A slight excess of the amine ensures complete forma- 303 K, interaction time 20, 40, 60, 90, 120, 180, 240, 300,
tion of the complex. The solution is diluted with water to 1 L to 360 min, pH 5.5 (Cd), 5.8 (Co), 5.7 (Cu, Pb, Ni).
give a 0.05 M solution of the complex. A quantity of 0.5 g of
dry adsorbent was mixed with 5 ml of the complex solution in 2.7.3. Effects of adsorbent amount
a 100-ml flask and diluted with distilled water to 25 ml and the Clay 2, 3, 4, 5, 6 g/L, metal ion concentration 50 mg/L,
mixture was agitated for 30 min in a thermostatic water bath temperature 303 K, interaction time 240 min (Cd, Co), 360 min
shaker and centrifuged. The concentration of the complex re- (Cu), 180 min (Pb, Ni), pH 5.5 (Cd), 5.8 (Co), 5.7 (Cu, Pb, Ni).
maining in the supernatant was determined by mixing 5 ml of
it with 5 ml of 0.1 M HCl to destroy the [Cu(en)2 ]2+ complex, 2.7.4. Effects of adsorbate concentration
followed by adding 0.5 g KI per ml and then titrating iodomet- Clay 2 g/L, metal ion concentration 10, 20, 30, 40, 50 mg/L,
rically with 0.02 M Na2 S2 O3 in the presence of starch as an temperature 303 K, interaction time 240 min (Cd, Co), 360 min
indicator. The CEC was calculated from the formula (Cu), 180 min (Pb, Ni), pH 5.5 (Cd), 5.8 (Co), 5.7 (Cu, Pb, Ni).
CEC (meq/100 g) = MSV (x − y)/1000m, (2) 2.7.5. Adsorption isotherm
where M is the molar mass of the complex, S the concentration Clay 2 g/L, metal ion concentration 10, 20, 30, 40, 50, 75,
of the thio solution, V the volume (ml) of the complex taken 100, 150, 200, 250 mg/L, temperature 303 K, interaction time
for iodometric titration, m the mass of adsorbent taken (g), 240 min (Cd, Co), 360 min (Cu), 180 min (Pb, Ni), pH 5.5 (Cd),
x the volume (ml) of thio required for blank titration (without 5.8 (Co), 5.7 (Cu, Pb, Ni).
the adsorbent), and y the volume (ml) of thio required for the
titration (with the adsorbent). 2.7.6. Adsorption thermodynamics
Clay 2 g/L, metal ion concentration 10, 20, 30, 40, 50 mg/L,
2.6. Metal adsorbate temperature 303, 308, 313 K, interaction time 240 min (Cd,
Co), 360 min (Cu), 180 min (Pb, Ni), pH 5.5 (Cd), 5.8 (Co),
The adsorption experiments were conducted using water 5.7 (Cu, Pb, Ni).
spiked with Cd(II), Co(II), Cu(II), Pb(II), and Ni(II), sep-
arately. Stock solutions containing 1000 mg of metal ion 2.8. Theoretical basis
per liter were prepared by dissolving appropriate amounts of
Cd(NO3 )2 ·4H2 O (Qualigens, Mumbai, India), Co(NO3 )2 ·6H2 O Two isotherm equations were used to find out the relation
(E. Merck, Mumbai, India), CuSO4 ·5H2 O (Qualigens, Mum- between the equilibrium concentrations of the adsorbate in the
bai, India), Pb(NO3 )2 (Glaxo, Mumbai, India), and Ni(NO3 )2 · liquid phase and in the solid phase. These isotherms [33,34] are
6H2 O (Qualigens, Mumbai, India) separately in double-distilled as follows:
water and was used to prepare the adsorbate solutions of the
required concentrations. The pH of the aqueous solution of (a) Freundlich isotherm:
Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) as prepared was 5.5,
qe = Kf Cen , (3)
5.8, 5.7, 5.7, and 5.7, respectively.
where Ce and qe are the equilibrium concentration of metal
2.7. Adsorption experiments ions in the liquid phase and in the solid phase, respectively,
Kf and n being Freundlich coefficients.
Batch adsorption experiments were carried out in 100-ml (b) Langmuir isotherm:
Erlenmeyer flasks by mixing together a constant amount of ad-
sorbent with a constant volume of the aqueous solution of metal Ce /qe = 1/(bqm ) + (1/qm )Ce , (4)
ions. The contents in the flasks were agitated by placing them where b and qm are Langmuir coefficients representing the
in a water bath thermostat for a known time interval. The mix- equilibrium constant for the adsorbate–adsorbent equilib-
ture was then centrifuged and the concentration of metal ions rium and the monolayer capacity. The linear Freundlich and
remaining unadsorbed in the supernatant liquid was determined Langmuir plots are obtained by plotting (i) log qe vs log Ce
by atomic absorption spectrometry (Varian SpectrAA 220 with and (ii) Ce /qe vs Ce respectively, from which the adsorp-
air–acetylene oxidizing flame). The pH of the adsorptive solu- tion coefficients could be evaluated.
tion was adjusted by addition of 0.01 N NaOH or 0.01 N HNO3
as needed. The following conditions were maintained for dif- The adsorption kinetics was tested with respect to the fol-
ferent sets of experiments: lowing well-known models:

2.7.1. Effect of pH (a) Pseudo-first-order kinetics using the Lagergren equation


Clay 2 g/L, metal ion concentration 50 mg/L, tempera- [35,36]:
ture 303 K, interaction time 240 min (Cd, Co), 360 min (Cu),
180 min (Pb, Ni), pH 1.0–10.0 (Cd, Co, Ni), 1.0–6.0 (Cu, Pb). ln(qe − qt ) = ln qe − k1 t, (5)
414 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

where qe and qt are the values of amount adsorbed per unit 3. Results and discussion
mass at equilibrium and at any time t, and k1 is the pseudo-
first-order adsorption rate constant. The values of k1 can be 3.1. Characterization of clay adsorbents
obtained from the slope of the linear plot of log(qe − qt )
vs t . 3.1.1. XRD study
(b) The second-order kinetics is applied when it is found that Acid treatment and calcination introduce some changes into
ln qe is not equal to the intercept of the first-order plot as the crystal structure of clay minerals. The intensity of the char-
obtained from Eq. (5) by using the equation [37] acteristic XRD peaks of the clay minerals are reduced on acid-
ification and the peaks become widened [41]. This implies
 
t/qt = 1/ k2 qe2 + (1/qe ) · t, (6) changes in the regular pattern of the clay structure along with
partial destruction compared to the original. The dispersion and
where k2 qe2 is described as the initial adsorption rate as amorphization of the acid-treated clay minerals are known to
t → 0. The plot of t/qt vs t gives a straight line, which give rise to an increase in the very-low-angle diffraction inten-
allows computation of qe and k2 . sities [42] and a reduction in the corresponding tip widths [43].
(c) When there is a possibility of diffusion of adsorbate species The XRDs of both the clay adsorbents are given in Fig. 1. The
into the interior of the pores of the adsorbent, the intraparti- following observations can be drawn from the XRD measure-
cle diffusion [38] rate constant (ki ) is obtained from the ments:
equation:
(i) For the calcined, acid-activated montmorillonite, the basal
spacing expanded from 4.44 to 4.47 Å (2θ = 19.98◦ ),
qt = ki · t 0.5 . (7) which was accompanied by a decrease in intensity from
43.06 to 29.09%.
Intraparticle diffusion plays a significant role in controlling
(ii) The intensity of most of the XRD peaks of montmoril-
the kinetics of the adsorption process, if the plots of qt vs lonite decreased sharply on acid treatment, so that both
t 0.5 yield straight lines passing through the origin with the octahedral and tetrahedral sites might have been affected
slope giving the rate constant, ki . drastically.
(d) The liquid film diffusion model [39] given by (iii) Acid activation of montmorillonite yielded two new peaks
at 22.91 Å (2θ = 3.85◦ ) and 12.49 Å (2θ = 7.06◦ ), which
ln(1 − F ) = −kfd t (8) were absent in the untreated montmorillonite. Appearance
of new peaks indicated the formation of expansible phases
is applicable when the flow of the reactant from the bulk and interlamellar expansion [44].
liquid to the surface of the adsorbent determines the rate (iv) The tip width of the 19.98◦ peak (2θ ) in montmorillonite
constant F and the fractional attainment of equilibrium decreased from 0.32 to 0.30 (acid-activated montmoril-
(qt /qe ) and kfd is the adsorption rate constant. A linear plot lonite).
of − ln(1 − F ) vs t with zero intercept suggests that the ki-
netics of the adsorption process is controlled by diffusion The XRD measurements showed that activation with 0.25 M
through the liquid film surrounding the solid adsorbent. H2 SO4 influenced the structural properties of montmorillonite.

The thermodynamic parameters for the adsorption process, 3.1.2. FTIR study
H (kJ/mol), S (J/(K mol)), and G (kJ/mol), are evalu- XRD results have already demonstrated some significant
ated using the equations [40] changes brought about by modification of montmorillonite
through acid treatment. The FTIR measurements have lent addi-
G = −RT ln Kd , (9) tional support where assignment of bands to different structural
features is done as per Farmer [45]. The FTIRs of the clays are
ln Kd = −G/RT = S/R − H /RT , (10) given in Fig. 2. This is summarized below:

where K d = the distribution coefficient of the adsorbate (= qe / (i) The number of OH-stretching bands is the same for both
Ce in L/g), T = the temperature (K), and R = the universal montmorillonite (M1) and acid-activated montmorillonite
gas constant (8.314 × 10−3 kJ/(K mol)). The plot of ln Kd vs (M2) with very little deviation of the band positions.
1/T should be linear with the slope (−H /R) and the in- However, all the OH-stretching bands have reduced in-
tercept (S/R) giving the values of H and S. The Gibbs tensities after acid treatment (3512 cm−1 : 76.5 to 65.4%;
energy change, G, is the fundamental criterion of spontane- 3554 cm−1 : 77.3 to 66.6%; 3620 cm−1 : 74.8 to 64.8%
ity. Processes occur spontaneously at a given temperature if G with the band shifting to 3621 cm−1 , etc.).
is a negative quantity. All these relations are valid when the en- (ii) The SiO-stretching band (1033 cm−1 ) for M2 occurs at the
thalpy change remains constant in the range of temperatures same position as that of M1, but the intensity decreases
considered. from 59.3 to 49.6%.
K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424 415

(a)
(a)

(b) (b)
Fig. 1. XRD of (a) montmorillonite M1 and (b) acid-activated montmorillonite Fig. 2. FTIR spectra of (a) montmorillonite M1 and (b) acid-activated montmo-
M2. rillonite M2.

(iii) The SiO-bending band at 464 cm−1 decreases in intensity OH groups of layered aluminosilicates and adsorbed water) and
from 72.1% (M1) to 61.1% after acid activation. The band (ii) 400–1150 cm−1 (due to lattice vibrations). Ravichandran
at 420 cm−1 (71.4%) shifts to 428 cm−1 with a decrease and Sivasankar [48] have demonstrated that the IR bands of
in intensity to 64.0%. montmorillonite and vermiculite activated with HCl did not ex-
(iv) The band at 722 cm−1 due to the presence of quartz in hibit any significant change from those of the parent clays. The
M1 did not change its position in M2; however, the band results in the present work are similar, but although principal IR
showed a decrease in intensity from 69.8 (M1) to 60.5% frequencies show only minor shifts, the intensities of the bands
(M2). are modified significantly. This has indicated that acid treatment
brings about considerable interaction between H+ ions and the
Montmorillonite has two characteristic FTIR regions [46– clay surface. It is also to be inferred that the interaction is taking
48], viz., (i) 3500–3750 cm−1 (due to the surface structural place with those surface groups whose IR frequencies exhibit a
416 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

maximum shift. The SiO-bending frequencies show maximum


shifts (∼8 cm−1 ), while the OH-stretching, OH-bending, and
SiO-stretching bands have not shifted much due to acid treat-
ment.

3.1.3. Surface area


The specific surface areas of montmorillonite and acid-acti-
vated montmorillonite were measured as 19.8 and 52.3 m2 /g,
respectively. Ravichandran and Sivasankar [48] reported a spe-
cific surface area of 19.0 m2 /g for montmorillonite, which on
treatment with HCl (0.1–0.7 M) increased to 188.3 m2 /g. Such
high values of specific surface area were not achieved in the
present work by treatment with 0.25 M H2 SO4 . The acid treat-
ment opens up the edges of the platelets and as a consequence,
the surface area and the pore diameter increase [43], which is
in conformity with the results obtained in this work. Kara et
al. [49] reported that an increase in the surface area of sepiolite (a)
upon acid activation followed by calcination could be attributed
to the removal of water molecules, both those formed during
acid activation and those inherently present as crystal water.

3.1.4. Cation exchange capacity (CEC)


The CEC of montmorillonite was measured as 153.0 meq/
100 g in agreement with the reported values [50]. On acid treat-
ment, the CEC increases to 341.0 meq/100 g.
Brønsted and Lewis acid sites are associated with the inter-
lamellar region and the edge sites of clays respectively. The ion-
exchange capacity of clay minerals is attributed to structural
defects, broken bonds, and structural hydroxyl transfers [51].
Treatment of the clay with 0.25 M H2 SO4 results in replace-
ment of a number of different cations with H+ ions and on
subsequent heating and calcination, dehydroxylation occurs,
leaving behind a number of Lewis sites. Much of the increase
in CEC in the present work is likely to be due to increase in
Lewis acidity, as the acid-treated clay was calcined at 773 K (b)
before CEC measurement. The CEC increase may be a com-
plex process involving these effects as well.

3.2. Adsorption of metal ions

3.2.1. Effects of pH
It is not possible to carry out adsorption experiments with
Cu(II) and Pb(II) at pH >6.0, or for Ni(II) and Co(II) at pH
>8.0, due to precipitation of the metals as the hydroxides,
which introduces uncertainty into the interpretation of the re-
sults. The precipitation was very rapid for Cu(II) and Pb(II) at
pH >6.0 and the experiments had to be terminated at pH 6.0.
Adsorption of Cd(II) could be studied over the pH range from
1.0 to 10.0 without any abrupt changes in adsorption. The gen-
eral trend in the extent of adsorption (%) as well as amount
(c)
adsorbed (qe ) with increase in pH is positive. Fig. 3 explains
the trend in the amount of metal ions adsorbed (qe ) with re-
Fig. 3. Influence of pH on the amount of metal ions adsorbed for adsorp-
spect to pH. The acid activation increases the number of sites tion of Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) on montmorillonite, M1 and
responsible for adsorption of various metal species and at any acid-activated montmorillonite, M2: (a) Cu(II) and Pb(II), (b) Co(II) and Ni(II),
pH, the acid-activated montmorillonite has a higher adsorption (c) Cd(II) at equilibrium time (clay 2 g/L, initial metal concentration 50 mg/L,
capacity than the nonactivated clay. temperature 303 K).
K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424 417

Table 1
Effect of interaction time for adsorption of Cd(II), Co(II), Cu(II), Pb(II), and
Ni(II) on montmorillonite (M1) and acid-activated montmorillonite (M2) at the
natural pH (clay 2 g/L, initial metal ion concentration 50 mg/L, temperature
303 K)
Adsorbate Clay ad- Equilibrium Extent of adsorption qe
sorbent time (%) (mg/g)
(min) 40 min Equilibrium 40 min Equilibrium
Cd(II) M1 240 79.2 86.2 19.8 21.6
M2 84.8 92.2 21.2 23.0

Co(II) M1 240 50.4 63.1 12.6 15.8


M2 58.0 69.5 14.5 17.4

Cu(II) M1 360 74.9 83.9 18.7 21.0


M2 78.3 87.5 19.6 21.9

Pb(II) M1 180 77.9 86.9 19.5 21.7


M2 84.5 90.8 21.1 22.7

Ni(II) M1 180 50.9 62.8 12.7 15.7


M2 57.4 69.2 14.4 17.3

At very low pH, the number of H3 O+ ions exceeds that


of metal ions by several times and the metal ions can hardly (a) (b)
compete with H3 O+ ions for the binding sites on the clay adsor-
bents. With an increase in pH, the concentration of H3 O+ ions Fig. 4. Lagergren pseudo-first-order plots for adsorption of Cd(II), Co(II),
Cu(II), Ni(II), and Pb(II) on (a) montmorillonite M1 and (b) acid-activated
decreases and some of the sites become available to the metal
montmorillonite M2 at the natural pH of the metal solution (clay 2 g/L, ini-
ions. As the acidity decreases, more and more H3 O+ ions on tial metal ion concentration 50 mg/L, temperature 303 K).
the clay surface are replaced by metal ions and metal species
(such as Cd2+ , Cd(OH)+ , and Cd(OH)2 [52]; Co2+ , Co(OH)+ ,
Table 2
Co(OH)2 , Co(OH)3 , and Co(OH)4 [53]; Cu2+ , Cu(OH)+ , and The first-order and second-order rate coefficients for adsorption of Cd(II),
+ −
Cu2 (OH)2+ 2+
2 [54]; Pb , Pb(OH) , Pb(OH)2 , Pb(OH)3 , and Co(II), Cu(II), Pb(II), and Ni(II) on montmorillonite (M1) and acid-activated
2− 2+ +
Pb(OH)4 [55]; and Ni and NiOH [56]). This happens at a montmorillonite (M2) at the natural pH (clay 2 g/L, initial metal ion con-
centration 50 mg/L, temperature 303 K, units of k1 and k2 are min−1 and
comparatively lower pH (∼6.0) for Cu(II) [57] and Pb(II) [58],
g/(mg min), respectively)
while the adsorption could be safely carried out to pH ∼10.0
Metal Clay Pseudo-first order Pseudo-second order
for Cd(II) [59] and ∼8.0 for Ni(II) [60] and Co(II) [49].
ion adsorbent k1 × 102 r k2 × 102 r
3.2.2. Effects of interaction time and kinetics of adsorption Cd(II) M1 1.9 −0.74 3.1 +0.99
M2 2.4 −0.99 11.1 +1.00
For all the metal ions, adsorption is very fast initially and
the maximum uptake is recorded within 40 min, which slowly Co(II) M1 2.7 −0.92 4.7 +0.99
approaches equilibrium. Table 1 gives details of the increase in M2 2.6 −0.93 5.4 +0.99
the extent of adsorption (%) as well as the amount adsorbed Cu(II) M1 2.8 −0.98 15.4 +0.99
per unit mass (qe ) for all the metal ions at their respective equi- M2 3.1 −0.98 15.9 +0.99
libria. The initial uptake rate for all the metals is very high, as Pb(II) M1 2.8 −0.99 8.4 +0.99
a large number of adsorption sites are available for adsorption M2 4.0 −0.97 11.3 +0.99
at the onset of the process. As the sites are gradually filled up,
Ni(II) M1 3.1 −0.97 5.3 +0.99
adsorption becomes slow and the kinetics becomes more de- M2 3.2 −0.96 5.5 +0.99
pendent on the rate at which the adsorptive is transported from
the bulk phase to the actual adsorption sites [61].
The rates of adsorption have been tested first with the pseu- plots does not establish a first-order mechanism for the interac-
do-first-order mechanism of Lagergren by plotting log(qe − qt ) tions [62]. This is observed when qe values obtained from the
vs t (min) (Fig. 4). The regression coefficients (r ∼ −0.74 to Lagergren plots are compared with the experimental qe values
−0.99) and the first-order rate constants obtained from these (Table 3).
plots (values from 1.9 × 10−2 to 3.1 × 10−2 min−1 for mont- The first-order kinetics is therefore inappropriate to ex-
morillonite M1 and 2.4 × 10−2 to 4.0 × 10−2 min−1 for acid- plain the rate processes. Second-order kinetic plots (t/qt vs t )
activated montmorillonite M2) are given in Table 2. All the (Fig. 5) had better linearity (r ∼ +0.99, Table 2) and appeared
plots are linear, showing that the interactions might be follow- to give a better understanding of the interactions. Acid acti-
ing a first-order mechanism and that the rate is dependent on vation enhances the second-order rate constant for adsorption
the initial concentration. However, linearity of the Lagergren of the metal ions for all the cases (values from 3.1 × 10−2
418 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

(a) (b)

(a) (b)
Fig. 6. Intraparticle diffusion plots for adsorption of Cd(II), Co(II), Cu(II),
Ni(II), and Pb(II) on (a) montmorillonite M1 and (b) acid-activated montmoril-
Fig. 5. Second-order plots for adsorption of Cd(II), Co(II), Cu(II), Ni(II), and lonite M2 at the natural pH of the metal solution (clay 2 g/L, initial metal ion
Pb(II) on (a) montmorillonite M1 and (b) acid-activated montmorillonite M2 at concentration 50 mg/L, temperature 303 K).
the natural pH of the metal solution (clay 2 g/L, initial metal ion concentration
50 mg/L, temperature 303 K).
Table 4
The intraparticle and liquid film diffusion rate coefficients for adsorption of
Table 3 Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) on montmorillonite (M1) and acid-
Experimental and computed qe values from Lagergren and second-order plots activated montmorillonite (M2) at the natural pH (clay 2 g/L, initial metal
for adsorption of Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) on montmorillonite ion concentration 50 mg/L, temperature 303 K, units of ki and kfd are mg/
(M1) and acid-activated montmorillonite (M2) at the natural pH (clay 2 g/L, (g min0.5 ) and min−1 , respectively)
initial metal ion concentration 50 mg/L, temperature 303 K) Metal Clay Intraparticle diffusion Liquid film diffusion
Metal Clay qe (mg/g) ion adsorbent ki × 10 Intercept r kfd × 102 Intercept r
ion adsorbents Experi- Lagergren Second Deviation (%)
Cd(II) M1 2.1 +18.5 +0.88 1.9 +1.8 +0.99
mental order Lagergren Second order M2 2.7 +19.4 +0.90 2.4 +1.6 +0.99
Cd(II) M1 21.6 3.5 18.1 −83.8 −16.2 Co(II) M1 4.3 +9.8 +0.96 2.6 +0.4 +0.92
M2 23.0 4.7 23.4 −79.5 +1.7 M2 4.0 +11.9 +0.96 1.8 +1.0 +0.99
Co(II) M1 15.8 11.9 16.6 −24.6 +5.0 Cu(II) M1 2.1 +17.9 +0.81 2.8 +1.6 +0.98
M2 17.4 10.2 18.1 −41.2 +4.0 M2 2.1 +18.8 +0.80 3.1 +0.2 +0.98
Cu(II) M1 21.0 4.1 21.3 −77.1 +1.2 Pb(II) M1 3.3 +17.3 +0.93 2.8 +1.1 +0.99
M2 21.9 4.4 22.2 −80.4 −1.3 M2 2.5 +19.4 +0.91 3.8 +0.9 +0.98
Pb(II) M1 21.7 7.4 22.3 −65.8 +2.7 Ni(II) M1 4.2 +10.1 +0.95 3.0 +0.3 +0.97
M2 22.7 10.2 23.2 −55.0 +2.2 M2 4.2 +11.6 +0.95 2.5 +0.7 +0.97
Ni(II) M1 15.7 12.7 16.5 −19.1 +5.0
M2 17.3 12.9 18.1 −25.4 +4.6
The plots (qt vs t 0.5 ) for intraparticle diffusion similarly
yield linear curves (Fig. 6, r ∼ +0.80 to +0.96) and the val-
to 15.4 × 10−2 g/(mg min) for montmorillonite M1 and 5.4 × ues of the rate constant, ki , are given in Table 4. Significantly,
10−2 to 15.9 × 10−2 g/(mg min) for acid-activated montmo- the plots do not have a zero intercept (range: +9.8 to +18.5
rillonite M2). The experimental qe values and those obtained for montmorillonite M1 and +11.6 to +19.4 for acid-activated
from the slopes of the second-order plots now show much bet- montmorillonite M2), indicating that the diffusion of metal
ter agreement (Table 3). Some deviations still exist that might species into the pores is not the dominating factor controlling
be due to the actual process not being in conformity with simple the mechanism of the process. Similar results are obtained with
first-order or second-order kinetics. the liquid film diffusion model, involving diffusion from the
K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424 419

creasing adsorbent amount provides greater surface area or ad-


sorption sites [63]. However, the amount adsorbed per unit mass
(qe ) decreases for both the clay adsorbents. A large amount of
adsorbent effectively reduces the unsaturation of the adsorption
sites and, correspondingly, the number of such sites per unit
mass comes down, resulting in comparatively less adsorption at
higher adsorbent amount.
For example, the amount of Cd(II) adsorbed per unit mass of
clay decreased from 21.6 to 8.2 mg/g and 23.0 to 8.3 mg/g for
montmorillonite and acid-activated montmorillonite, respec-
tively, as the clay amount was changed from 2.0 to 6.0 g/L.
The corresponding extent of adsorption increased from 86.2 to
98.5% and from 92.2 to 99.9% in the same order. Similar trends
were observed in the cases of clay–Co(II), clay–Cu(II), clay–
Pb(II), and clay–Ni(II) systems. The variation in the extent of
adsorption (%) and the amount adsorbed per unit mass (qe ) for
all clay–metal systems is given in Table 5. Similar observations
have been reported earlier (Pb(II) on sawdust [64], Cd(II) on
river bed sediment [65], etc.).
(a) (b)
3.2.4. Effects of initial metal ion concentration
Fig. 7. Liquid film diffusion plots for adsorption of Cd(II), Co(II), Cu(II), Ni(II), The number of metal ions adsorbed per unit mass (qe ) of
and Pb(II) on (a) montmorillonite M1 and (b) acid-activated montmorillonite clay increases gradually with increase in metal concentration,
M2 at the natural pH of the metal solution (clay 2 g/L, initial metal ion con- whereas the extent of adsorption (%) decreases with increasing
centration 50 mg/L, temperature 303 K). metal ion loading (Table 6). For example, by changing the ini-
tial concentration of Cd(II) from 10 to 50 mg/L, qe increases
bulk liquid phase to the surface of the adsorbent. The plots of from 4.7 to 21.6 mg/g and 5.0 to 23.0 mg/g for montmoril-
− ln(1 − F ) vs t are linear (Fig. 7, r ∼ +0.92 to +0.99) with lonite (M1) and acid-activated montmorillonite (M2), respec-
intercepts of +0.3 to +1.8 for montmorillonite M1 and +0.2 tively, whereas the extent of adsorption (%) decreases from 94.9
to +1.6 for acid-activated montmorillonite M2 (Table 4). The to 86.2% and from 99.7 to 92.2%, respectively, for M1 and M2.
model requires the curves to pass through the origin and al- Similar trends have been found in all other cases.
though the intercepts are not zero, their small values signify At low initial metal ion concentration, the ratio of the num-
some role played by diffusion from the liquid phase to the ad- ber of metal ions to the number of available adsorption sites
sorbent surface. The film diffusion rate coefficients are given in is small and consequently the adsorption is independent of the
Table 4. initial concentration, but as the concentration of metal ions in-
creases, the situation changes and the competition for adsorp-
3.2.3. Effects of adsorbent amount tion sites becomes fierce [66]. At high concentration of metal,
In all the cases, the extent of adsorption (%) increases a unit mass of the adsorbent is exposed to a larger number of
rapidly with increase in the adsorbent amount. This is to be metal ions and a progressively higher number of metal ions are
expected because, for a fixed initial solute concentration, in- taken up with the gradual filling up of the appropriate bind-

Table 5
Effect of clay amount on extent of adsorption (%) and amount of metal ion adsorbed per unit mass of clays (qe , mg/g) for adsorption of Cd(II), Co(II), Cu(II),
Pb(II), and Ni(II) on montmorillonite (M1) and acid-activated montmorillonite (M2) at the natural pH (initial metal concentration 50 mg/L, temperature 303 K) at
clay–metal equilibrium
Clay Clay amount Cd(II) Co(II) Cu(II) Pb(II) Ni(II)
adsorbent (g/L) % qe % qe % qe % qe % qe
M1 2 86.2 21.6 63.1 15.8 83.9 21.0 86.9 21.7 62.8 15.7
3 91.0 15.2 68.4 11.4 87.8 14.6 91.5 15.3 67.9 11.3
4 95.2 11.9 73.5 9.2 91.4 11.4 95.4 11.9 72.4 9.1
5 98.2 9.8 77.8 7.8 94.7 9.5 98.9 9.9 76.2 7.6
6 98.5 8.2 78.7 6.6 97.7 8.1 99.1 8.3 78.3 6.5

M2 2 92.2 23.0 69.5 17.4 87.5 21.9 90.8 22.7 69.2 17.3
3 95.2 15.9 74.7 12.5 91.4 15.2 94.0 15.7 74.5 12.4
4 97.7 12.2 79.6 10.0 94.9 11.8 97.2 12.3 79.3 9.9
5 99.6 10.0 83.8 8.4 98.0 9.8 99.2 9.9 83.4 8.3
6 99.9 8.3 84.8 7.1 99.3 8.3 99.9 8.3 85.5 7.1
420 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

Table 6
Effect of initial concentration of adsorbates on extent of adsorption (%) and amount of metal ion adsorbed per unit mass of clays (qe ) for adsorption of Cd(II),
Co(II), Cu(II), Pb(II), and Ni(II) on montmorillonite (M1) and acid-activated montmorillonite (M2) at the natural pH (clay 2 g/L, temperature 303 K) at clay–metal
equilibrium
Clay Metal ions Cd(II) Co(II) Cu(II) Pb(II) Ni(II)
adsorbent (mg/L) % qe % qe % qe % qe % qe
M1 10 94.9 4.7 84.6 4.2 96.6 4.8 95.3 4.8 84.3 4.2
20 92.2 9.2 78.1 7.8 92.5 9.3 92.3 9.2 77.8 7.8
30 89.5 13.4 71.9 10.8 88.9 13.3 90.4 13.6 71.6 10.7
40 87.9 17.6 67.0 13.4 89.0 17.2 88.7 17.7 66.7 13.3
50 86.2 21.6 63.1 15.8 83.9 21.0 86.9 21.7 62.8 15.7
M2 10 99.7 5.0 91.8 4.6 99.9 5.0 99.9 5.0 91.5 4.6
20 98.1 9.8 85.0 8.5 96.1 9.6 98.1 9.8 84.7 8.5
30 95.4 14.3 78.7 11.8 92.5 13.9 95.8 14.4 78.4 11.7
40 93.9 18.8 73.5 14.7 89.6 17.9 92.0 18.4 73.2 14.6
50 92.2 23.0 69.5 17.4 87.5 21.9 90.8 22.7 69.2 17.3

(a) (b)
(a) (b)
Fig. 9. Freundlich isotherm for Cd(II), Co(II), Cu(II), Ni(II), and Pb(II) on (a)
Fig. 8. Isotherm plots (qe vs Ce ) for adsorption of Cd(II), Co(II), Cu(II), Ni(II), montmorillonite M1 and (b) acid-activated montmorillonite M2 at the natural
and Pb(II) on (a) montmorillonite M1 and (b) acid-activated montmorillonite pH of the metal solution (clay 2 g/L, initial metal ion concentration 10, 20,
M2 at the natural pH of the metal solution (clay 2 g/L, initial metal ion con- 30, 40, 50, 75, 100, 150, 200, 250 mg/L, temperature 303 K) at clay–metal
centration 10, 20, 30, 40, 50, 75, 100, 150, 200, 250 mg/L, temperature 303 K) equilibrium.
at clay–metal equilibrium.

lonite and acid-activated montmorillonite have lower uptake of


ing sites. This gives rise to an increase in qe , although the Ni(II) and Co(II) than of Cd(II), Cu(II), and Pb(II).
percent adsorption comes down. Similar results have been re- The empirical Freundlich isotherm yields linear plots (r ∼
ported earlier for metal ion adsorption on different adsorbents +0.93 to +0.98, Fig. 9) and the values of the coefficients, n
(e.g., Cd(II) on activated carbon from coconut coir pith [67]; (all values <1.0) and Kf (4.5 to 9.2 mg1−1/n L1/n /g for mont-
Cu(II) on maple sawdust [61]; Ni(II) and Cd(II) on bagasse fly morillonite and 5.9 to 12.9 mg1−1/n L1/n /g for acid-activated
ash [68], etc.). montmorillonite, Table 7) indicate that the clays have a good
potential to be used as adsorbents for the metal ions Cd(II),
3.2.5. Adsorption isotherm Co(II), Cu(II), Ni(II), and Pb(II). It is to be noted that the
For initial metal ion concentrations of 10, 20, 30, 40, 50, Freundlich isotherm applies to adsorption on nonspecific and
75, 100, 150, 200, and 250 mg/L, the plots of qe vs Ce for heterogeneous sites on solid surfaces, but no definite mecha-
metal–clay interactions for a fixed amount of adsorbent (2 g/L) nism could be arrived at. The isotherm is valid for weak van
at 303 K are presented in Fig. 8. It is seen that both montmoril- der Waals type adsorption as well as for strong chemisorption.
K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424 421

Table 7
Freundlich and Langmuir coefficients for adsorption of Cd(II), Co(II), Cu(II),
Pb(II), and Ni(II) on montmorillonite (M1) and acid-activated montmorillonite
(M2) at the natural pH (clay 2 g/L, initial metal concentration 10, 20, 30, 40,
50, 75, 100, 150, 200, 250 mg/L, temperature 303 K, units of Kf , qm and b are
mg 1−1/n L1/n /g, mg/g and L/g, respectively) at clay–metal equilibrium
Metal Clay Freundlich coefficients Langmuir coefficients
ion adsorbent Kf n r qm b r
Cd(II) M1 8.6 0.3 +0.93 32.7 191.7 +0.99
M2 12.9 0.2 +0.97 33.2 225.4 +0.99

Co(II) M1 4.6 0.4 +0.97 28.6 72.1 +0.99


M2 6.0 0.3 +0.98 29.7 100.7 +0.99

Cu(II) M1 9.2 0.3 +0.97 31.8 192.8 +0.99


M2 12.4 0.2 +0.98 32.3 278.0 +0.99

Pb(II) M1 8.9 0.3 +0.93 33.0 206.6 +0.99


M2 11.3 0.3 +0.97 34.0 320.3 +0.99

Ni(II) M1 4.5 0.4 +0.97 28.4 71.5 +0.99


M2 5.9 0.3 +0.98 29.5 100.2 +0.99

The Kf values of acid-activated montmorillonite are higher than (a) (b)


those of the nonactivated montmorillonite. Fig. 10. Langmuir isotherm for adsorption of Cd(II), Co(II), Cu(II), Ni(II), and
The Langmuir isotherm is specific to strong monolayer Pb(II) on (a) montmorillonite M1 and (b) acid-activated montmorillonite M2 at
chemisorption, which might be the case in the present work the natural pH of the metal solution (clay 2 g/L, initial metal ion concentra-
since the Langmuir plots have good linearity (Fig. 10, r ∼ tion 10, 20, 30, 40, 50, 75, 100, 150, 200, 250 mg/L, temperature 303 K) at
clay–metal equilibrium.
+0.99) (Table 7). The Langmuir equilibrium coefficient, b, de-
termines the direction in which the adsorbate–adsorbent equi-
librium and Langmuir monolayer capacity qm = 32.68 to 39.22 mg/g,
whereas the uptake of Pb(II) on the electric furnace slag in the
clay(solid phase) + metal(II)(aqueous phase) = clay–metal(II) same temperature range had Freundlich and Langmuir mono-
is shifted. Large values of b ensure that the equilibrium is pre- layer capacity in the range from 0.57 to 1.82 mg1−1/n L1/n /g
dominantly driven toward the right, leading to the formation of and 33.78 to 37.04 mg/g, respectively [70]. Lopez et al. [71] re-
the adsorbate–adsorbent complex. ported that the Langmuir adsorption capacity for Ni(II) on red
From the values of Langmuir monolayer capacity, qm , it is mud was found as 10.95 mg/g.
observed that acid activation did not have much influence on the
values. This suggests that treatment with acid does not increase 3.2.6. Adsorption thermodynamics
the number of adsorption sites to a large extent, although the Adsorption of metal ions on clay adsorbents shows different
treatment influences the strength of the existing sites as revealed behavior with respect to increase in temperature (Table 8).
by the equilibrium coefficient (b) data. Langmuir monolayer Cu(II) adsorption follows an endothermic path. The extent
capacity, qm , is considerable for both the clays and therefore the of adsorption (%) and the amount adsorbed per unit mass of
clays can take up large amounts of the metal ions. The qm values the adsorbents (qe ) increase appreciably in the temperature
for Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) on montmorillonite range from 303 to 313 K. Cu(II) uptake changes from 83.9%
(M1) are 32.7, 28.6, 31.8, 33.0, and 28.4 mg/g and those for (21.0 mg/g) to 89.3% (22.3 mg/g), and 87.5% (21.9 mg/g)
acid-activated montmorillonite (M2) are 33.2, 29.7, 32.3, 34.0, to 92.7% (23.2 mg/g) for montmorillonite (M1) and acid-
and 29.5 mg/g in the same order of metal ions. activated montmorillonite (M2), respectively. Thus, adsorption
Similar ranges of adsorption coefficients for metal ions on of Cu(II) onto clays has to overcome a small activation energy
various adsorbents are reported in the literature. Adsorption of barrier and increasing energy supply makes it easier for Cu(II)
Cd(II) on perlite follows the Freundlich isotherm model with to adsorb onto the clay surface. Such activated adsorption fol-
an adsorption capacity of 0.64 mg/g [59]. Yuvuz et al. [25] lowing an endothermic path has also been reported earlier
have reported a Langmuir monolayer capacity, qm , of 0.919 [72,73]. The effect of increasing temperature is quite different
and 1.470 mg/g at 298 and 313 K, respectively, for Co(II) ad- in the case of Cd(II), where increase of temperature from 303
sorption on Turkish kaolinite. Langmuir adsorption capacity for to 308 K is found to enhance the amount of Cd(II) adsorbed per
Cu(II) adsorption on natural clinoptilolite and synthetic zeo- unit mass of the clays (as well as extent of adsorption, %), but
lites has been shown to be 5.91 and 50.5 mg/g, respectively, by the trend is reversed between 308 and 313 K. For example, in
Alvarez-Ayuso et al. [69]. The uptake of Cu(II) on electric fur- the case of montmorillonite, the adsorption of Cd(II) increases
nace slag in the temperature range from 293 to 313 K has Fre- from 21.6 mg/g (86.2%) to 22.8 mg/g (91.1%) on increasing
undlich adsorption capacity Kf = 0.44 to 1.09 mg1−1/n L1/n /g the temperature from 303 to 308 K, but the adsorption decreases
422 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

Table 8
Effect of temperature on extent of adsorption (%) and amount of metal ion adsorbed per unit mass of clays (qe , mg/g) for adsorption of Cd(II), Co(II), Cu(II),
Pb(II), and Ni(II) on montmorillonite (M1) and acid-activated montmorillonite (M2) at the natural pH (clay 2 g/L, initial concentration of metal ions 50 mg/L) at
clay–metal equilibrium
Clay Temperature Cd(II) Co(II) Cu(II) Pb(II) Ni(II)
adsorbent (K) % qe % qe % qe % qe % qe
M1 303 86.2 21.6 63.1 15.8 83.9 21.0 86.9 21.7 62.8 15.7
308 91.1 22.8 62.0 15.5 86.9 21.7 79.0 19.8 56.8 14.2
313 89.7 22.4 59.8 15.0 89.3 22.3 74.7 18.7 51.3 12.8
M2 303 92.2 23.0 69.5 17.4 87.5 21.9 90.8 22.7 69.2 17.3
308 93.1 23.3 68.3 17.1 89.9 22.5 84.3 21.1 61.5 15.4
313 92.6 23.2 66.1 16.5 92.7 23.2 79.1 19.8 55.5 13.9

Table 9 now almost in the region of chemisorption enthalpies. The


Thermodynamic data for adsorption of Cd(II), Co(II), Cu(II), Pb(II), and Ni(II) trends are similar for exothermic adsorption of Ni(II) (mont-
on montmorillonite (M1) and acid-activated montmorillonite (M2) at the nat- morillonite and acid-activated montmorillonite: H −45.2 and
ural pH (clay 2 g/L, initial concentration of metal ions 50 mg/L) at clay–metal
equilibrium
−56.9 kJ/mol) and Co(II) (montmorillonite and acid-activated
montmorillonite: H −15.1 and −19.9 kJ/mol). The mag-
Metal Clay H S −G (kJ/mol)
ions adsorbent (kJ/mol) (J/(K mol))
nitude of the adsorption enthalpy, H , indicates moderately
303 K 308 K 313 K Mean
strong bonding between the metal ions and the clay minerals.
Cd(II) M1 +40.2 +147.5 44.6 45.4 46.1 45.4 The endothermic enthalpy change (H ) associated with
M2 +25.0 +107.3 32.5 33.0 33.6 33.0
Cu(II) adsorption decreases on acid activation (montmoril-
Co(II) M1 −15.1 −47.1 14.3 14.5 14.7 14.5 lonite and acid-activated montmorillonite: H +50.7 and
M2 −19.9 −59.3 17.9 18.2 18.5 18.2 +45.8 kJ/mol). Similar endothermic interaction has been ob-
Cu(II) M1 +50.7 +180.7 54.7 55.6 55.9 55.4 served in Cd(II) adsorption (montmorillonite and acid-activated
M2 +45.8 +334.5 101.3 102.8 104.6 102.9 montmorillonite: H +40.2 and +25.0).
Pb(II) M1 −75.5 −235.9 71.4 72.5 73.8 72.5 Decrease in entropy during Co(II), Pb(II), and Ni(II) adsorp-
M2 −83.4 −258.7 78.3 79.5 80.8 79.5 tion helps to stabilize the clay–metal adsorption complex (S
are in the range from −47.1 to −235.9 J/(K mol) for mont-
Ni(II) M1 −45.2 −146.4 44.4 45.1 45.8 45.1
M2 −56.9 −181.8 55.0 55.8 56.8 55.8 morillonite and −59.3 to −258.7 J/(K mol) for acid-activated
montmorillonite). S values for interactions of Cu(II) and
Cd(II) with montmorillonite are +180.7 and +147.5 J/(K mol)
to 22.4 mg/g (89.7%) on further increase of temperature to and those for acid-activated montmorillonite are +334.5 and
313 K. The process is controlled by the adsorbate–adsorbent +107.3 J/(K mol), respectively. The endothermic interactions
and adsorbate–adsorbate forces and it is clear from the results are driven by entropy increase, which illustrates increased ran-
that the first one becomes weak in comparison to the latter as domness at the solid–solution interface indicating strong affin-
the temperature increases. ity of the adsorbent for Cu(II) and Cd(II).
For Co(II), Ni(II), and Pb(II), increase in temperature results The clay–metal interactions take place spontaneously and
in a decrease in the extent of adsorption (%) and the amount are accompanied by decrease in Gibbs energy (Table 9).
of metal adsorbed per unit mass of clay (qe ), showing the in- G values vary from −14.5 to −72.5 kJ/mol for montmo-
teractions to be exothermic. For example, in the temperature rillonite and from −18.2 to −102.9 kJ/mol for acid-activated
range of 303 to 313 K, Co(II) uptake changes from 63.1% montmorillonite, respectively.
(15.8 mg/g) to 59.8% (15.0 mg/g) and 69.5% (17.4 mg/g) Thermodynamic data on metal adsorption on clays are
to 66.1% (16.5 mg/g) for montmorillonite and acid-activated scarce. Yuvaz et al. [25] reported the thermodynamic para-
montmorillonite, respectively. Thus, the adsorbate ions have a meters for adsorption of Cu(II), Ni(II), Co(II), and Mn(II) on
tendency to escape from the solid phase to the solution. More- Turkish kaolinite. H and S values were +39.52 kJ/mol
over, rise in temperature and excess energy supply promote and +11.7 J/(K mol) for Cu(II), +37.27 kJ/mol and +10.7 J/
desorption. Such results are not uncommon for adsorption of (K mol) for Ni(II), +21.52 kJ/mol and +5.4 J/(K mol) for
metal ions from the aqueous phase [74]. Co(II), and +36.73 kJ/mol and +10.1 J/(K mol) for Mn(II).
The thermodynamic adsorption parameters, H , S, and Echeverria et al. [75] have reported that H , S, and G for
G (Table 9), are computed from the plots of ln Kd vs 1/T . adsorption of Ni(II) onto illite have values of +16.8 kJ/mol,
In the case of clay–Pb(II) interaction, the mean adsorption +58 J/(mol K), and −1.04 kJ/mol, respectively. Lin and
enthalpy, H , is −75.5 kJ/mol for montmorillonite, which Juang [28] investigated the adsorption of Cu(II) and Zn(II)
changes to −83.4 kJ/mol after acid activation. The exothermic onto montmorillonite modified with sodium dodecyl sulfate.
enthalpy is higher after acid activation, with a clear indica- The values of H , S, and G were reported as 7.05 kJ/mol,
tion that Pb(II) ions are held more strongly by acid-activated 9.09 J/(K mol), and −9.66 kJ/mol, respectively, for Cu(II) and
montmorillonite. H for acid-activated montmorillonite is 7.39 kJ/mol, 6.39 J/(K mol), and −9.17 kJ/mol, respectively,
K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424 423

for Zn(II) at 298 K. These values are not much different from [13] D.C.K. Ko, C.W. Cheung, K.K.H. Choy, J.F. Porter, G. McKay, Chemo-
the values obtained in this work. sphere 54 (2004) 273.
[14] C.R.T. Tarley, M.A.Z. Arruda, Chemosphere 54 (2004) 987.
[15] V.K. Gupta, I. Ali, J. Colloid Interface Sci. 271 (2004) 321.
4. Conclusions [16] V.K. Gupta, P. Singh, N. Rahman, J. Colloid Interface Sci. 275 (2004) 398.
[17] S. Maity, S. Chakravarty, S. Bhattacharjee, B.C. Roy, Water Res. 39
Montmorillonite and its acid-activated forms are capable of (2005) 2579.
removing metal ions, namely Cd(II), Co(II), Cu(II), Ni(II), and [18] U. Wingenfelder, B. Nowack, G. Furrer, R. Schulin, Water Res. 39 (2005)
Pb(II), from aqueous solution. Acid activation enhances the ad- 3287.
[19] V.K. Gupta, A. Rastogi, V.K. Saini, N. Jain, J. Colloid Interface Sci. 296
sorption capacity compared to the untreated clay mineral due to
(2006) 59.
increased surface area and pore volume. Adsorption increases [20] K. Tanabe, in: J.R. Anderson, M. Boudart (Eds.), Catalysis—Science and
with pH till the metal ions are precipitated out in alkaline solu- Technology, Springer-Verlag, New York, 1981.
tion. [21] R. Naseem, S.S. Tahir, Water Res. 35 (2001) 3982.
The rate of uptake of the metal ions by the clays is very [22] A. Mellah, S. Chegrouche, Water Res. 31 (1997) 621.
high initially, followed by a low rate indicating the entry of [23] S.A. Khan, R. Rehman, M.A. Khan, Waste Manage. 15 (1995) 271.
[24] A.G. Sanchez, E.A. Ayuso, O.J. De Blas, Clay Miner. 34 (1999) 469.
the metal ions into the interior of the adsorbent particles as
[25] O. Yuvaz, Y. Altunkayank, F. Guzel, Water Res. 37 (2003) 948.
the interactions proceed. The processes are much more akin to [26] E. Alvarez-Ayuso, A. Garcia-Sanchez, Clays Clay Miner. 51 (2003) 475.
second-order kinetics. The uptake of metal ions depends appre- [27] D.G. Strawn, N.E. Palmer, L.J. Furnare, C. Goodell, J.E. Amonette, R.K.
ciably on the amount of clay. The amount adsorbed per unit Kukkadapu, Clays Clay Miner. 52 (2004) 321.
mass (qe ) and the net adsorption (%) show reverse trends with [28] S.-H. Lin, R.-S. Juang, J. Hazard. Mater. B 92 (2002) 315.
increase in adsorbent amount. [29] K.G. Bhattacharyya, S. Sen Gupta, Ind. Eng. Chem. Res. 45 (2006) 7232.
[30] A.G. Espantaleon, J.A. Nieto, M. Fernandez, A. Marsal, Appl. Clay Sci.
Both Langmuir and Freundlich isotherms yield good fits 24 (2003) 105.
and the adsorption coefficients agree well with the condi- [31] G.W. Sears, Anal. Chem. 28 (1956) 1981.
tions supporting favorable adsorption. All the metal–clay in- [32] F. Bergaya, M. Vayer, Appl. Clay Sci. 12 (1997) 275.
teractions are thermodynamically favorable. Clay–Pb(II), clay– [33] H.M.F. Freundlich, J. Phys. Chem. 57 (1906) 385.
Ni(II), and clay–Co(II) interactions are exothermic in nature, [34] I. Langmuir, J. Am. Chem. Soc. 40 (1918) 1361.
[35] S. Lagergren, Kungl. Svenska Vetensk. Handl. 24 (1898) 1.
whereas clay–Cu(II) and clay–Cd(II) are endothermic. The
[36] Y.S. Ho, Scientometrics 59 (2004) 171.
clay–metal interactions are spontaneous in all cases supported [37] Y.-S. Ho, C.-C. Chiang, Y.-C. Hsu, Sep. Sci. Technol. 36 (2001) 2473.
by decrease in Gibbs energy. [38] W.J. Weber, J.C. Morris, J. Sanit. Engg. Div. Am. Soc. Civ. Eng. 89 (1963)
31.
Acknowledgment [39] G.E. Boyd, A.W. Adamson, L.S. Myers Jr., J. Am. Chem. Soc. 69 (1947)
2836.
[40] W.J. Thomas, B. Crittenden, in: Adsorption Technology and Design,
One of the authors (S.S.G.) is grateful to the University Butterworths–Heinemann, Oxford, 1998, chap. 1.
Grants Commission, New Delhi, for providing assistance for [41] M. Radojevc, V. Jovic, D. Karaulic, D. Vitorovic, J. Serb. Chem. Soc. 67
this work under the FIP scheme. (2002) 499.
[42] G. Jozefaciuk, G. Bowanko, Clays Clay Miner. 50 (2002) 771.
References [43] F.R.V. Diaz, P.D.S. Santos, Quim. Nova 24 (2001) 345.
[44] D.S. Fanning, V.Z. Keramidas, M.A. El-Desosky, in: J.B. Dixon, S.B.
Weed (Eds.), Minerals in Soil Environments, in: SSSA Book Series, vol. 1,
[1] ATSDR, Toxicological profile for cadmium, Agency for Toxic Substances
Soil Science Society of America, Medison, WI, 1989.
and Disease Registry (http://www.atsdr.cdc.gov), U.S. Department of
Health and Human Services, Atlanta, GA, 1999. [45] V.C. Farmer, in: H. van Olphen, J.J. Fripiat (Eds.), Data Handbook for
[2] ATSDR, Draft toxicological profile for cobalt, Agency for Toxic Sub- Clay Materials and Other Non-Metallic Minerals, Pergamon Press, Ox-
stances and Disease Registry (http://www.atsdr.cdc.gov), U.S. Department ford, 1979.
of Health and Human Services, Atlanta, GA, 2001. [46] M.L. Hair, Infrared Spectroscopy in Surface Chemistry, Marcel Dekker,
[3] ATSDR, Draft toxicological profile for copper, Agency for Toxic Sub- New York, 1967.
stances and Disease Registry (http://www.atsdr.cdc.gov), U.S. Department [47] H. Suquet, S. Chevalier, C. Marcilly, D. Barthomeuf, Clay Miner. 26
of Health and Human Services, Atlanta, GA, 2002. (1991) 49.
[4] ATSDR, Toxicological profile for lead, Agency for Toxic Substances and [48] J. Ravichandran, B. Sivasankar, Clays Clay Miner. 45 (1997) 854.
Disease Registry (http://www.atsdr.cdc.gov), U.S. Department of Health [49] M. Kara, H. Yuzer, E. Sabah, M.S. Celik, Water Res. 37 (2003) 224.
and Human Services, Atlanta, GA, 1999. [50] R.E. Grim, Clay Mineralogy, McGraw Hill, New York, 1968.
[5] ATSDR, Draft Toxicological profile for nickel, Agency for Toxic Sub- [51] M.G.F. Rodrigues, Ceramica 49 (2003) 146.
stances and Disease Registry (http://www.atsdr.cdc.gov), U.S. Department [52] C.-H. Lai, C.-Y. Chen, B.-L. Wei, S.-H. Yeh, Water Res. 36 (2002) 4943.
of Health and Human Services, Atlanta, GA, 2003. [53] Y.-H. Wang, S.-H. Lin, R.-S. Juang, J. Hazard. Mater. B 102 (2003) 291.
[6] H.H. Tran, F.A. Roddick, J.A. O’Donnell, Water Res. 33 (1999) 2992. [54] S.T. Bosso, J. Enzweiler, Water Res. 36 (2002) 4795.
[7] V.K. Gupta, M. Gupta, S. Sharma, Water Res. 35 (2001) 1125. [55] V.C. Taty-Costodes, H. Fauduet, C. Porte, A. Delacroix, J. Hazard. Mater.
[8] R.L. Ramos, J.L.A. Bernal, B.J. Mendoza, R.L. Fuentes, C.R.M. Guerrero, B 105 (2003) 121.
J. Hazard. Mater. B 90 (2002) 27. [56] B. Bayat, J. Hazard. Mater. B 95 (2002) 251.
[9] Y.S. Ho, J.F. Porter, G. McKay, Water Air Soil Pollut. 141 (2002) 1. [57] Y. Nuhoglu, E. Oguz, Process Biochem. 38 (2003) 1627.
[10] S.-C. Pan, C.-C. Lin, D.-H. Tseng, Resour. Conserv. Recycl. 39 (2003) 79. [58] C.K. Jain, D. Ram, Water Res. 31 (1997) 154.
[11] L.J. Yu, S.S. Shukla, K.L. Dorris, A. Shukla, J.L. Margrave, J. Hazard. [59] T. Mathialagan, T. Viraraghavan, J. Hazard. Mater. 94 (2002) 291.
Mater. B 100 (2003) 53. [60] V. Padmavathy, P. Vasudevan, S.C. Dhingra, Process Biochem. 38 (2003)
[12] N. Chiron, R. Guilet, E. Deydier, Water Res. 37 (2003) 3079. 1389.
424 K.G. Bhattacharyya, S.S. Gupta / Journal of Colloid and Interface Science 310 (2007) 411–424

[61] B. Yu, Y. Zhang, A. Shukla, S.S. Shukla, K.L. Dorris, J. Hazard. Mater. [69] E. Alvarez-Ayuso, A. Garcia-Sanchez, X. Querol, Water Res. 37 (2003)
B 80 (2000) 33. 4855.
[62] Y.S. Ho, G. McKay, Adsorp. Sci. Technol. 20 (2002) 797. [70] L. Curkovic, S. Cerjan-Stefanovic, A. Rastovean-Mioe, Water Res. 35
[63] S. Rengaraj, K.-H. Yeon, S.-Y. Kang, J.-U. Lee, K.-W. Kim, S.-H. Moon, (2001) 3436.
J. Hazard. Mater. B 92 (2002) 185. [71] E. Lopez, B. Soto, M. Arias, A. Nunez, D. Rubinos, M.T. Barral, Water
[64] B. Yu, Y. Zhang, A. Shukla, S.S. Shukla, K.L. Dorris, J. Hazard. Mater. Res. 32 (1998) 1314.
B 84 (2001) 83. [72] A. Shukla, Y.-H. Zhang, P. Dubey, J.L. Margrave, S.S. Shukla, J. Hazard.
[65] C.K. Jain, M.K. Sharma, Water Air Soil Pollut. 137 (2002) 1. Mater. 95 (2002) 137.
[66] H. Ucun, Y.K. Bayhan, Y. Kaya, A. Cakici, O.F. Algurb, Desalination 154 [73] Y.-S. Ho, Water Res. 37 (2003) 2323.
(2003) 233. [74] M.M. Abou-Mesalam, Colloids Surf. A Physicochem. Eng. Aspects 225
[67] K. Kadirvalu, C. Namasivayam, Adv. Environ. Res. 7 (2003) 471. (2003) 85.
[68] V.K. Gupta, C.K. Jain, I. Ali, M. Sharma, V.K. Saini, Water Res. 37 (2003) [75] J. Echeverria, J. Indurain, E. Churio, J. Garrido, Colloids Surf. A Physic-
4038. ochem. Eng. Aspects 218 (2003) 175.

You might also like