You are on page 1of 10

Available online at www.sciencedirect.

com

Applied Clay Science 42 (2008) 214 – 223


www.elsevier.com/locate/clay

Synthesis and characterization of tannin-immobilized hydrotalcite as a


potential adsorbent of heavy metal ions in effluent treatments
Thayyath S. Anirudhan ⁎, Padmajan S. Suchithra
Department of Chemistry, University of Kerala, Kariavattom, Trivandrum-695 581, India
Received 1 August 2007; received in revised form 10 December 2007; accepted 16 December 2007
Available online 28 December 2007

Abstract

The modification of calcined hydrotalcite (HTC) by immobilizing tannin was investigated in the present work, in order to apply the modified media
(TA-HTC) as an adsorbent in the removal of heavy metal ions (Cu(II), Zn(II) and Cd(II)) from industrial effluents. The adsorbent was characterized with
FTIR, XRD, SEM, TG/DTG, surface area analyzer and potentiometric titrations. Batch experiments were performed as a function of process parameters
such as agitation time, initial metal concentration, pH, ionic strength and adsorbent dose. The maximum adsorption was found at pH 6.0. The mechanism
for the removal of metal ions by TA-HTC was based on ion exchange process. Experimental results showed that the adsorption of these metal ions was
selective to be in the order of Cu(II) N Zn(II) N Cd(II). The process was very fast initially and maximum adsorption was observed with 3 h of agitation. The
adsorption kinetics were investigated and kinetic parameters such as rate constant, external mass transfer diffusion and intraparticular mass transfer
diffusion coefficients were evaluated. The adsorption process obeyed the intraparticular diffusion model. An increase of ionic strength of the medium
caused a decrease in metal ion adsorption. Equilibrium isotherm data were analyzed by the Langmuir, Freundlich and Dubinin–Radushkevich equations
using non-linear regression analysis. The best interpretation for the equilibrium data was given by the Langmuir isotherm and the maximum adsorption
capacities were 81.47 mg g− 1 for Cu (II), 78.91 mg g− 1 for Zn(II) and 74.97 mg g− 1 for Cd(II). The adsorption efficiency towards heavy metal ion removal
was tested using different industry wastewaters. The adsorption efficiency, regenerative and reuse capacities of this adsorbent were also assessed for four
consecutive adsorption/ desorption cycles and were found to retain the adsorption capacity.
© 2008 Elsevier B.V. All rights reserved.

Keywords: Tannin; Hydrotalcite; Heavy metal ions; Adsorption isotherm; Wastewater; Desorption

1. Introduction molecules. HTs are generally used for the removal of anions from
aqueous solutions and wastewaters. On heating at the range 500–
The use of hydrotalcites, HTs (also known as layered double 800 °C, HTs produce magnesium and aluminium oxides solid
hydroxides or anionic clays) for the purification of water is well solutions. The calcined product, HTC (Mg1 − x AlxO1 + x/2) has a
understood and progress today is mainly in improving the effi- memory effect and therefore it is possible to reconstruct the
ciency and specificity of the HTs and their applications. The original HT from aqueous solutions containing anions as
general formula of HTs is [M12+− x Mx3+ (OH)2]x+ [Ax/nn−mH2O]x−, expressed by Eqs. (1) and (2) (Martin et al., 1999):
where M2+ and M3+ are divalent and trivalent metal ions such as
Mg2+ and Al3+, An− is an n-valent anions and x can have values ½Mg1−x Alx ðOHÞ2 ðCO3 Þx=2 ➝Mg1−x Alx O1þx=2
between 0.20 and 0.33. In this structure, some of the M2+ ions are þ ðx=2ÞCO2 þ H2 O ð1Þ
replaced by M3+ ions, resulting in positively charged layers where
by the charge is compensated by interlayer anions. Water mole- Mg1−x Alx O1þx=2 þ ðx=nÞAn − þ ½1 þ ðx=2Þ
cules are present in the interlayer spaces. These anions and water þ yH2 O➝½Mg1−x Alx ðOHÞ2 ðAn − Þx =n þ xOH− ð2Þ
molecules can be interchanged with other anions and polar

⁎ Corresponding author. Tel.: +91 471 2418782. The ability of HTC to recover its structure provides the
E-mail address: tsani@redifflmail.com (T.S. Anirudhan). potential for reuse and recycle of the adsorbent, which is important
0169-1317/$ - see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.clay.2007.12.002
T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223 215

for their applications. Recently, there are increasing interests in 2.2. Preparation of hydrotalcites
using HTCs, because of their high anion exchange capacity and
surface area, for removal of inorganic oxyanions such as chromate The carbonate form of Mg–Al HT with Mg/Al molar ratio equal to 3 was
synthesized by the co-precipitation method at constant pH. An aqueous solution
(Manju et al., 1999), selenite (You et al., 2001), phosphate (Seida (250 mL) containing 0.75 mol of Mg(NO3)2 6H2O and 0.25 mol of Al(NO3)3
and Nakeno, 2002), arsenate and arsenite (Yang et al., 2005) and 9H2O was added drop wise under vigorous mechanical stirring with 500 mL
organic species such as surfactants (Pavlovic et al., 1997), phe- solution containing 1.7 mol of NaOH and 0.5 mol of Na2CO3. The resulting
nolic compounds (Barriga et al., 2002) and pesticides (Pavlovic slurry was heated to 70 °C for 24 h followed by filtration and washing with
et al., 2005). Surface modification of HTC can be used as a deionised water and was then dried at 120 °C. The chemical analysis (AAS) of
technique for the removal of cations from aqueous systems. To the resulting HT indicates that its Mg/Al ratio is 2.91 (Anirudhan and Suchithra,
2007). To prepare the calcined form (HTC), HT was heated from room
adsorb cations, the modified surface must either possess nega- temperature to 500 °C for 2 h. The HTC was cooled to 30 °C and placed in a
tively charged exchange sites, or there should be replacement of desiccator.
weakly held counter ions of the modifier by more strongly held
adsorbate counter ions. There are few reports on the adsorption of 2.3. Preparation of tannin-immobilized HTC
chelating agents such as ethylenediamminetetraacetate (EDTA)
and nitirlotri acetate (NTA) on HTC to form EDTA/NTA-HTC Tannin was loaded to HTC at pH 4.0 by the equilibrium adsorption of tannin
complexes and application of these complexes as adsorbents for from aqueous solution followed by drying in air at 70 °C. The pH of the solution
for tannin adsorption was selected on the light of our earlier adsorption
metal cations (Tarasov et al., 2003; Pevez et al., 2006). Since experiments that clarified that the maximum tannin removal was occurred in the
tannin contains hydroxyl functional groups, it was employed as a initial pH range 3.0–5.0 (Anirudhan and Suchithra, 2007). A significant decline
modifier to improve the surface properties and adsorption capacity in tannin adsorption was occurred above and below this pH range. The
of the activated carbon towards the metal cations uptake (Ucer et maximum adsorption at pH range 3.0–5.0 is due to the phenolic hydroxyl
al., 2006). groups of tannin and the hydrogen-bonding sites on the HTC. When tannin
molecule comes near reaction sites on the HTC surface, participation in the
Our recent work on use of calcined and uncalcined HTs as
reconstruction of the HTC according to the following reaction is favored.
adsorbents has demonstrated high adsorption potential for tannin
from aqueous solutions and coir industry effluents (Anirudhan and Mg3 AlO4 ðOHÞ þ 4H2 O þ TA− ➝½Mg3 AlðOHÞ8 TA þ OH− ð3Þ
Suchithra, 2007). This study showed that tannin adsorption on HT The decrease in the adsorption at a higher pH is possibly due to the increased
or HTC is not completely reversible and the bonding between the solubility of tannin and the abundance of OH− ions there by increasing
adsorbent particle and adsorbed tannin is likely strong. The use of hindrance to diffusion of tannate anions.
It was possible to design an operational condition for loading a maximum
an adsorbent in wastewater treatment depends not only in the
amount of tannin onto HTC based on the adsorption isotherm. Adsorption
adsorption capacity, but also on how well the spent adsorbent can isotherm was obtained by shaking 0.1 g of HTC with 50 mL solution of tannin of
be regenerated and used again. In many cases after four or five the desired concentrations (100–3000 μmol L− 1) at pH 4.0 for 24 h to attain
cycles of repeated use, the adsorbent materials can not be reused satisfactory equilibrium. The tannin concentrations in the supernatant solutions
and causing a disposal problem. The disposal of the spent adsor- were determined by means of UV–visible spectrophotometer (λmax 270 nm).
The uptake of tannin was calculated from the decrease of tannin concentrations
bent in an economically sound manner is very important. Any
relative to that of initial concentrations. The tannin immobilized HTC was
attempt to reutilize the spent adsorbent will be worthwhile. To the referred to as TA-HTC. The adsorbent was filtered and washed with distilled
best of our knowledge, no attempt has been done to examine the water to remove unadsorbed tannin from the adsorbent surface and then dried at
applicability of the spent HTs as adsorbent for water purification. 70 °C in air. The adsorbent with an average particle size of 0.096 mm was used
On continuity to our program, the present work was decided to for adsorption experiments.
investigate the possible application of the tannin adsorbed HTC as
adsorbent for the subsequent adsorption of heavy metals from 2.4. Adsorption measurements
aqueous solutions. Batch experiments were conducted to inves- Batch equilibrium method was used to investigate the adsorption of Cu(II),
tigate the adsorption potential of Cu(II), Zn(II) and Cd(II) ions Zn(II) and Cd(II) ions onto TA-HTC. Adsorption isotherms were obtained using
from aqueous solutions by tannin-immobilized HTC. solutions containing just one metal ion at a time, using exactly the same
conditions for the three cations. Fifty ml of metal solution containing 100 mg of
2. Materials and methods the adsorbent in a 100 ml Erlenmeyer flask was agitated at 200 rpm in a water
bath, where temperature was controlled at 30 °C. The pH of the solution was
2.1. Materials maintained at a definite value (6.0) using 0.1 M HCl or 0.1 M NaOH solutions.
Ten different concentrations of metal ions between 10–400 mg L− 1 were used.
All chemicals are reagent grade and were used as received with out further After attaining equilibrium, the aqueous phase was separated by centrifugation
purification. All solutions were prepared with deionised water. Metal solutions and the final concentration of metal ions in the solution was determined by AAS.
were prepared by dissolving appropriate amount of CuCl2 2H2O, CdCl2 2.5H2O The adsorbed amount of metal ions was calculated from the difference between
and ZnCl2 (Fluka, Switzerland) in distilled water. 0.1 M NaOH and HCl the initial and final concentrations.
solutions were used for pH adjustment. A Systronic microprocessor pH meter For each kinetic experiment, 50 ml aqueous solution of known concentration
(model μ-362) was used for pH measurements. A UV–visible spectro- was taken in the flask and shaken in a water bath at 30 °C. A 100 mg of the
photometer (Jasco model V-530) was used for the determination of tannin in adsorbent was added into the flask and the kinetic measurements were started.
solution. A GBC Avanta (A 5450) atomic absorption spectrophotometer (AAS) At given time intervals, the solutions were centrifuged and the supernatants were
with copper, cadmium and zinc hollow cathod lamps and air acetylene flame analyzed for metal ions as mentioned above. The effect of solution pH was
was used for determining Cu(II), Cd(II) and Zn(II) ions in solution. A studied in the range 2.0–7.0. Initial pH values were adjusted using 0.1 M NaOH
temperature controlled water bath flask shaker (Lab line, India) was used for or 0.1 M HCl solution. Initial concentration of metal ions was 25 mg L− 1, and
shaking all the solutions. equilibrium time was 3 h.
216 T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223

2.5. Desorption and regeneration studies

The desorption studies were carried out with varying concentrations of HCl
solution. Hundred mg of TA-HTC was dispersed in 10 mg L− 1 of metal ions
solutions as mentioned in the adsorption studies. The metal ion loaded TA-HTC was
separated from the mixture by centrifugation and washed with deionised water to
remove any unadsorbed metal. To this 50 ml of 0.1 M HCl solution was added. The
flasks were shaken at 200 rpm for 3 h at 30 °C using water bath shaker. The
adsorbent was then separated by centrifugation and supernatant was analyzed for
the desorbed metal ion concentration in a similar way described previously. The
desorption ratio was calculated from the amount of metal ions adsorbed on the TA-
HTC and the final metal ion concentration in the adsorption medium. In order to
determine the reusability of the adsorbent, consecutive adsorption–desorption Fig. 1. qe versus Ce plot for the adsorption of tannin onto HTC.
cycles were repeated five times by using the same adsorbent.

2.6. Characterization methods optimum concentrations and pH range in order to further


investigate their adsorption characteristics for Cu(II), Cd(II) and
The specific surface area of the HT and HTC and TA-HTC was analyzed
using a Quantasorb (QS-7) surface area analyzer that uses a nitrogen Zn(II) ions.
adsorption–desorption method. The total pore volume was determined from
the amount of N2 adsorbed at a relative pressure of 0.95. The amount of carbon 3.2. Stability of the TA-HTC
in HT, HTC and TA-HTC was determined using a CHN analyzer (Carlo Erba-
1361). FTIR analysis of the adsorbent samples was performed on a Schimadzn
In order to find out whether there is any leakage of TA from
FTIR model 1801. The surface morphology of the adsorbent samples was
analyzed using a Phillips XC-3CP scanning electron microscope. A potentio- TA-HTC during the adsorption experiments, the supernatant
metric titration method [Schwarz et al., 1984] was used to determine the point of solution at different pH values were analyzed. It was observed
zero charge (pHpzc). The surface charge density σ0 of HT, HTC and TA-HTC that there was no leakage from the adsorbent in a pH range 3.0–
was determined by using the equation 8.0. However at a higher pH range 8.5–10.0, a small amount of
TA (5–11%) leached out from TA-HTC. Since the wastewater
F ðCA  CB Þ þ ½OH   ½Hþ 
r0 ¼ ð4Þ containing metal ions are always acidic in nature and the
A
optimum pH range for wastewater treatment is below 7.0, a
where F is the Faraday constant (96487 C eq− 1), CA and CB are the concen- small amount of TA leaching at higher pH may not create any
trations of the acid and base (eq L− 1) after each addition during the titration. H+
problem. The adsorption stability of the 0.1 g metal ion-loaded
and OH− ions are the equilibrium concentration of H+ and OH− ions bound to the
suspension surface (eq cm− 2) and A is the specific surface area of the suspension TA-HTC was also determined in 50 mL 0.1 M HCl solution. No
surface (cm2 L− 1). The point of intersection of σ0 with the pH curves gives the remarkable leaching of tannin (4–7%) from the spent adsorbent
pHpzc (pH at which σ0 is zero). was observed in the resulted solutions by UV–visible spectro-
photometry in four cycles. The process of adsorption and
3. Results and discussion desorption does not alter the original physical and chemical
characteristics of the adsorbent.
3.1. Loading capacity of tannin
3.3. Adsorbent characterization
The effect of pH on tannin adsorption was studied over the pH
range 2.0–9.0 by adjusting the pH value of 100 μmol L− 1 tannin The carbon contents of the HT, HTC and TA-HTC were 8.3,
solution with 0.1 M HCl or 0.1 M Na OH and then agitated 0.1 g 0.6 and 36.2% respectively. A comparison of these values reveals
HTC for 3 h. The results showed that the removal percentage that CO32− was almost removed at the end of the calcination and the
increased from 68.3 to 99.2% with raising pH from 2.0 to 3.0 and tannin molecules were exchanged in the structure. The HT
remained constant up to pH 5.0. Turning to alkaline range, it displays a point of zero charge at pHpzc = 7.6. After calcination, the
decreased to 33.0% at pH 9.0. The maximum adsorption was pHpzc of the HT has increased to 8.4. It is interesting to note that
achieved at pH range 3.0–5.0 and this may be due to the phenolic Chatelet et al. (1996) concluded, based on their studies on mea-
hydroxyl groups of tannin and the hydrogen–bonding sites on the surements of pHpzc of calcined and uncalcined HTs , that the value
HTC. of pHpzc, after calcinations, increased from 7.6 to 8.9 in 0.01 M
Loading capacity can be judged in a more systematic and Na2SO4 electrolyte, but the value decreased from 7.8 to 6.6 in 0.01
quantitative way through adsorption isotherm analysis. Fig. 1 M Na2CO3 and remained unchanged (11.1) in 0.01 M NaCl. After
shows the experimental tannin uptake isotherm for HTC at pH 4.0 calcination at 500–800 °C, HT produces Mg and Al oxides. These
and 30 °C. In the figure, the adsorption appears to take place oxides increase the acidic nature of the HTC (Arco et al., 2001)
according to Langmuir-type adsorption, i.e., the adsorption capa- and so the value of pHpzc was expected to be decreased. It is
city increased with the equilibrium concentration of tannin in unclear at this time why the value of pHpzc in the present study,
solution, progressively reaching saturation of the adsorbent. increased after calcination in 0.01 M NaCl. Additional work is
Langmuir adsorption capacity of HTC was calculated as high as presently being performed to evaluate the variation of surface
397.65 μmol g− 1. After knowing the maximum loading capacity of charge of HT and HTC as a function of pH in different electrolytes.
HTC, corresponding HTCs were loaded with tannin at their The increase in pHpzc after calcination indicates that the HTC
T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223 217

became more positive. After tannin loading, the pHpzc of HTC has
decreased to 6.0, at pH above that value the surface charge of the
clay has a negative charge, but below that pH the clay has a
positive charge. The BET surface area was calculated from N2
adsorption data and values were found to be 37.8 and 48.9 m2 g− 1
for HTC and TA-HTC, respectively. The total pore volume deter-
mined from the N2 adsorption data at P/P0 of 0.95 was found to be
0.41 and 0.47 mL g− 1 for HTC and TA-HTC, respectively.
The FTIR spectra of HT, HTC and TA-HTC are shown in
Fig. 2. The broad band observed in HT spectrum around
3490 cm− 1 is due to the vibration of structural OH− groups from
the brucite-like layers. The peak at about 1640 cm− 1 is assigned to
Fig. 3. XRD patterns of HT, HTC and TA-HTC.
HOH deformation (δH–O–H) confirming the presence of water
molecules in the interlayer region of HT. The bands observed at
approximately 1365, 875 and 670 cm− 1 are assigned to carbonate to those of natural hydrotalcite reported in the literature (7.6–
ions. The spectrum of HTC, suggests that most of the inter layer 7.8 Å). The disappearance of this characteristic basal reflection in
water bands have disappeared but some of the carbonate anions are HTC confirms that CO32− is removed from the interlayer with out
still present after calcination (e.g. band at ~1370 cm− 1). However causing any collapse in the crystal structure. It can be seen from
these may also be due to the carbonate species adsorbed when the Fig. 3 that the XRD reflections in the case of HTC are compa-
HTC sample is cooling in the atmosphere during calcination. The ratively small and broad. Fig. 3 shows XRD patterns of the TA-
FTIR spectrum of the TA-HTC shows band characteristics of the HTC with a diffraction peak of 300 verifying the presence of the
aliphatic CH stretching at 2920 cm− 1 and broad band in the region tannin molecule. After reaction with tannin, the basal spacing was
3500 cm− 1 due to phenolic OH stretching. Another feature in TA- 9.2 Å with increased intensity of the reflections. Loading of tannin
HTC spectrum is the bands at 1394 and 1620 cm− 1 attributed to the lead to a considerable increase of crystallinity in comparison to the
symmetrical and asymmetrical vibrations of COO−, the band parent HTC sample indicated by the increased number and
at 1617 cm− 1 is corresponding to the stretching C = C groups in intensity of the XRD reflections. The value of d110 = 1.53 Å for HT
aromatic rings. The FTIR spectrum of the Cu(II)-TA-HTC is also decreases slightly for HTC (d110 = 1.49 Å) and HTC-tannin
shown in Fig. 2. The peak at 1368 cm− 1 is indicative of C–O (d110 = 1.43 Å), which agrees with decrease of molar ratio Mg/Al.
stretching at the benzene ring when Cu(II) gets coordinated with
oxygen. The band at 517 cm− 1 represents Cu–O band.
The XRD patterns of HT, HTC and TA-HTC are presented in
Fig. 3, which illustrates the 003 and 006 reflections. The reflection
patterns in these samples were identified by the previous reports
(Ulibarri et al., 2001). It was observed that all the samples possess
reflection characteristics of layered structure with intense lines at
low 2θ values and less intense and asymmetric lines at higher 2θ
values. The characteristic basal spacing d003 = 7.57 Å, correspond-
ing to the interlayer CO32− anions observed in HT is almost similar

Fig. 2. FTIR spectra of HT, HTC, TA-HTC and Cu-TA-HTC. Fig. 4. SEM images obtained at 1500x magnification of HT, HTC and TA-HTC.
218 T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223

the reaction mixture varied between 3.0 and 6.0, the final pH of the
reaction medium remained between 2.6 and 4.4 for Cu(II), 2.6 and
4.7 for Cd(II) and 2.8 and 5.2 for Zn(II) at an initial concentration
of 25 mg L− 1. It is thus considered that in the pH range 3.0 – 6.0
metal cations are adsorbed by releasing protons from the phenolic
OH groups of TA present in TA-HTC, according to cation ex-
change mechanism. With increase of pH, the enhancement of
adsorption is apparently due to the hydrolysis of the exchanging
cations, since the hydroxyl complexes, that is M(OH)+ ions are
adsorbed in preference to the uncomplexed cations (Das and
Bandyopadhyay, 1992). At pH values higher than 6.0, both ion
exchange and aqueous metal hydroxide formation (not necessarily
Fig. 5. TG/DTG curves of HTC and TA-HTC. precipitation) may become significant mechanisms in the metal
removal process. Thus to correlate removal with ion exchange, an
Fig. 4 shows the morphology of the HT, HTC and TA-HTC optimum initial pH chosen for performing kinetic and isotherm
and provides a basis for comparison between the adsorbents. experiments was 6.0.
HT shows an adhesive and globule like surface and has patchy
type of particle distribution. After calcination, its form changed 3.5. Adsorption kinetics
to a corn flake-like structure. This corn flake appearance pro-
bably occurs due to the reduction in certain crystalline phase The adsorption of the studied metal ions onto TA-HTC as a
originally associated within the HT. The SEM photograph of function of contact time was investigated and data were given in
TA-HTC reveals that surface of the HTC was spread and Fig. 7. The adsorption was rapid in the first 10 min and then
covered with a cloudy like surface with widely scattered pores slowed considerably as the reaction approached equilibrium
due to the tannin molecules. after 3 h for all metals. The equilibrium time is independent of
The TG and DTG curves for HTC and TA-HTC are shown in initial concentration. The time profile of metal uptake is a
Fig. 5. In HTC there is three-stage decomposition and first de- single, smooth and continuous curve leading to saturation,
composition occurs at 90 °C due to dehydration, in which a weight suggesting the possible monolayer coverage of metal ion on the
loss of 9.0% is produced. The second decomposition temperature surface of the adsorbent. With increase in metal concentration
is 323 °C, with a weight loss of 19.0% due to dehydroxylation. In from 25 to 100 mg L− 1, the amount of metal adsorbed increased
the third stage at 680 °C, a weight loss of 36.0% is produced due to from 12.46 to 40.05 mg g− 1 for Cu(II), 12.31 to 38.50 mg g− 1
decarboxylation as well as dehydroxylation. In the case of TA- for Zn(II) and 12.12 to 35.65 mg g−1 for Cd(II), indicating the
HTC having two stages of decomposition, first occurs at 103 °C metal removal by adsorption on TA-HTC is concentration
with a weight loss of 10.3% due to dehydration. The second dependent.
decomposition of TA-HTC occurs at 392 °C, in which a weight
loss of 30.0% is produced by dehydroxylation and also by the 3.6. Kinetic models
decomposition of tannin. From these results it is clear that TA-
HTC has a better thermal stability. Quantifying the changes in adsorption with time requires an
appropriate kinetic model. In the present study we employed four
3.4. Effect of pH on metal ions adsorption on TA-HTC kinetic models namely, pseudo-first-order, Second-order Elovich
equation, modified Freundlich model and pseudo-second-order
The wastewater containing metal ions present commonly on (Arslanoglu et al., 2005). On comparison, it was found that re-
acidic pH. The effect of pH on Cu(II), Cd(II) and Zn(II) adsorption cently introduced pseudo-second-order kinetic model yielded
was studied in the range 2.0–7.0. The metal ion concentration and good agreement between the experimentally obtained kinetic
adsorbent dose were kept constant at 25 mg L− 1 and 2 g L− 1
respectively (Fig. 6). At lower pH values, the functional groups of
TA-HTC will be protonated to a higher extent and this results in a
strong repulsion to the positively charged ions in solution. For this
reason Cu(II), Cd(II) and Zn(II) ions were best adsorbed onto TA-
HTC at pH in the range 5.0–7.0. A maximum adsorption of 99.7%
Cu(II), 98.9% Zn(II) and 97.2% Cd(II) was observed at an initial
concentration of 25 mg L− 1 at pH 6.0. According to the simple
speciation diagrams which were constructed by Leyva-Ramos
et al. (1997), Baraket (2005), Gibert et al. (2005) for Cd(II), Cu(II)
and Zn(II), respectively, all the species occurring at pH values of
7.0 and below carry a positive charge either as Cd2+, Cd(OH)+,
Cu2+, Cu(OH)+, Zn2+ or Zn(OH)+ ions. It has been shown that the Fig. 6. Effect of pH on the adsorption of Cu(II), Zn(II) and Cd(II) ions onto
final pH is always less than the initial pH. When the initial pH of TA-HTC.
T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223 219

second-order rate constant. Plotting 1/q t against 1/t the


constants k2 and qe for different concentrations can be deduced
(Table 1). Comparison of the experimental results and the
calculated values of metal adsorption on solid phase by means
of Eq. (5) (in which the value of k2 and qe were inserted)
indicates that the adsorption of metals onto TA-HTC can be
best described by pseudo-second-order kinetic model. The
values of k2 decrease with an increase of initial concentration,
while the value of qe increase with the increase of initial
concentration. The lower the concentration of metal ion in
solution, the lower the probability of collision between these
ions and hence the faster these metal ions could be bonded to
the active sites on the adsorbent surface. The previously
mentioned kinetic model basically includes all steps of
adsorption such as external mass transfer diffusion adsorption
and intraparticular mass transfer diffusion, so that is pseudo-
model.

3.7. Mass transfer aspects of metal ions sorption onto TA-HTC

3.7.1. External mass transfer model


The model used in this study is based on Fickien's laws. This
model expresses the evaluation of the concentration of solute in
the solution, C (mg L− 1), as a function of the difference in the
concentration of the metal ions in the solution and the adsorbent
surface, Cs (mg L− 1), according to the equation of McKay and
Bino (1988)
dC=dt ¼ BL S ðC  Cs Þ ð6Þ
−1
where BL is the external mass transfer coefficient (cm s ) and
Fig. 7. Effect of contact time and initial concentration on the adsorption of S is the specific surface area for mass transfer. The coefficients
Cu(II), Zn(II) and Cd(II) ions onto TA-HTC. are determined after making some assumptions such as Cs is
negligible at t = 0, C tending to the initial concentration (C0) and
also negligible intraparticular diffusion. So Eq. (6) is simplified
profiles and those obtained from the kinetic model. The following to
modern formulation of the pseudo-second-order kinetic model
was used (Shibi and Anirudhan, 2005). ½d ðC=C0 Þ=dt tY0 ¼ BL S ð7Þ
 
1=qt ¼ 1= k2 q2e t þ 1=qe ð5Þ 3.7.2. Intraparticular mass transfer diffusion model
In the model developed by Urano and Tachikawa (1991), the
where qe and qt are the amounts of metal ions adsorbed at adsorption rate is considered as independent of the stirring
equilibrium and at time (t) respectively, k2 is the pseudo- speed and the external diffusion is negligible relative to the low

Table 1
Kinetic parameters for the adsorption of Cu(II), Zn(II) and Cd(II) ions onto TA-HTC
Metal ion Concentration (mg L− 1) k2 (g mg− 1 in− 1) qe, exp (mg g− 1) qe, cal (mg g− 1) Di (cm2 s− 1) BL (cm s− 1)
Cu(II) 25 3.95 × 10− 2 12.25 12.21 3.03 × 10− 13 2.15 × 10− 4
50 2.06 × 10− 2 23.75 23.07 2.58 × 10− 13 2.01 × 10− 4
75 1.87 × 10− 2 32.67 31.91 2.05 × 10− 13 1.85 × 10− 4
100 1.35 × 10−2 40.00 39.09 1.64 × 10− 13 1.74 × 10− 4
Zn(II) 25 3.18 × 10−2 12.13 12.09 2.69 × 10− 13 2.11 × 10−4
50 1.64 × 10− 2 23.25 22.66 2.22 × 10− 13 1.99 × 10− 4
75 1.55 × 10−2 31.13 30.08 1.81 × 10− 13 1.67 × 10− 4
100 1.31 × 10− 2 38.51 37.02 1.59 × 10− 13 1.56 × 10− 4
Cd(II) 25 2.94 × 10− 2 11.87 11.75 2.40 × 10− 13 1.98 × 10− 4
50 1.42 × 10− 2 22.82 21.84 1.75 × 10− 13 1.85 × 10− 4
75 1.08 × 10− 2 30.75 29.81 1.61 × 10− 13 1.62 × 10− 4
100 0.97 × 10−2 38.00 35.11 1.32 × 10− 13 1.00 × 10− 4
220 T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223

also leading to a decrease in adsorption with increase in ionic


strength. In the chloride medium, uncharged species (e.g.
CuCl2, ZnCl2 and CdCl2) and charged complexes (e.g. CuCl+,
ZnCl+ and CdCl+) will be formed. These chloro-species are less
strongly adsorbed than Cu2+, Zn2+ and Cd2+ ions because
they have lower affinities to TA-HTC surface. Therefore, the
coexistence of Cl−ions limits the adsorption of cations from
aqueous solutions containing NaCl electrolyte.

3.9. Adsorption isotherms

Fig. 8. Effect of ionic strength on the adsorption of Cu(II), Zn(II) and Cd(II) ions The experimental adsorption isotherms for metal ions onto
onto TA-HTC. TA-HTC from aqueous solutions were generated at 30 °C for
concentration ranges between 25 and 400 mg L− 1. (Fig. 9). The
capacity of the adsorbent for the removal of metal ions from
overall adsorption rate. The adsorption kinetics is modelised aqueous solutions is commonly represented by three adsorption
according to the following equation: models, Langmuir, Freundlich and Dubinin-Radushkevich. The
h  i Langmuir equation has been used extensively for dilute solu-
f ðqt =qe Þ ¼  log 1  ðqt =qe Þ2 ¼ 4p2 Di t=2:3d 2 ð8Þ tions in the following form

where Di is the diffusion coefficient in the solid (cm2 s− 1) and d qe ¼ Q0 bCe =ð1 þ bCe Þ ð9Þ
the particle diameter. The values of BL and Di for different
where Q0 (mg g− 1) is the amount of solute adsorbed per unit
concentrations were determined from the plots of C/C0 versus t
weight of adsorbent required for monolayer coverage of the
and log[1 − (qt/qe)2] versus t respectively (figures not shown)
surface and b the constant related to the heat of adsorption. The
and the results are presented in Table 1. Since increasing the
Langmuir model assumes monolayer adsorption on a surface
metal concentration in the solution reduced the diffusion of
with a finite number of identical sites, that all sites are energe-
metal ions in the boundary layer, the values of BL decreased
tically equivalent and that there is no interaction between ad-
with increasing C0. The values of BL gradually decreased, and
sorbed solutes.
this is the reason for decreasing the metal ion adsorption by TA-
HTC with the gradual increase in metal concentration. It is
noted that the values of BL presented here is in consistent with
the reported values in the literature (Jansson-Charrier et al.,
1996), suggesting that the velocity of transport of the adsorbate
from the liquid phase to solid phase is rapid enough to use TA-
HTC for the treatment of wastewater rich in metal ions.
According to the earlier workers (Michelson et al., 1975), the
values of Di in the range 10− 11–10− 13 cm2 s− 1 reflects pore
diffusion as the rate-limiting step. It is evident that the rate-
limiting step appears to be pore diffusion process since the
magnitude of the coefficient is in the range of 10− 13 cm2 s− 1.

3.8. Effect of ionic strength

The effect of ionic strength on metal ion adsorption was


studied by conducting batch experiments at different ionic
strength of 0.001, 0.005, 0.01, 0.05 and 0.1 M NaCl. The results
are shown in Fig. 8. It was observed that when the ionic strength
increased from 0.001 to 0.1 M, the percentage removal
decreased from 98.3 to 60.9% for Cu(II), 96.2 to 56.9% for
Zn(II) and 91.4 to 46.8% for Cd(II), at an initial concentration of
25 mg L− 1. The adverse effect of ionic strength on metal ion
uptake suggests the possibility of ion exchange mechanism
being in operation in the adsorption process. Na+ ions (present
in salt used to change the ionic strength of metal solution) can
compete with Cu2+, Zn2+ and Cd2+ ions for cation exchange Fig. 9. Comparison of the experimental and model fits of the Langmuir,
sites in the TA-HTC surface and thus negatively affect the Freundlich and DR Isotherms (lines) for the adsorption of Cu(II), Zn(II) and
removal efficiency. The modification of the metal species is Cd(II) ions onto TA-HTC.
T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223 221

Table 2
Isotherm parameters for the adsorption of metal ions onto TA-HTC
Metal Langmuir Freundlich DR
Q0 (mg g− 1) b (L mg− 1) r2 χ2 KF 1/n r2 χ2 χm (mg g− 1) β r2 χ2
Cu(II) 81.47 0.115 0.997 1.23 24.97 0.226 0.936 8.6 29.80 0.0061 0.981 9.4
Zn(II) 78.91 0.099 0.999 1.17 23.02 0.232 0.923 6.4 27.79 0.0064 0.921 7.8
Cd(II) 74.97 0.082 0.994 1.43 20.70 0.238 0.914 7.2 25.26 0.0070 0.901 8.6

The Freundlich expression is an empirical equation based on fore Cd2+ ions (the largest diameter) have the minimum adsorp-
adsorption onto a heterogeneous surface. The Freundlich equa- tion, and Cu2+ (the least diameter) have maximum adsorption.
tion has the form The numerical value of Freundlich constants 1/n b 1 for all
metals indicates that adsorption capacity is only slightly sup-
qe ¼ KF Ce1=n ð10Þ pressed at lower equilibrium concentrations. The ultimate adsorp-
where KF and 1/n are Freundlich constants, indicating adsorp- tion capacity of TA-HTC can be calculated from the isothermal
tion capacity and adsorption intensity, respectively. This iso- data using Freundlich equation. Thus for an equilibrium concen-
therm usually fits the experimental data over a wide range of tration of 1 mg L− 1 each gram of TA-HTC can remove 24.97,
concentrations. 23.02 and 20.70 mg (KF) of Cu(II), Zn(II) and Cd(II), respectively
The Dubinin–Radushkevich (DR) has been used to describe at 30 °C. The adsorption energy (E) can be calculated using DR
the adsorption of metals on clays. The DR equation has the form equation and the following relationship has been used.
pffiffiffiffiffiffiffiffiffi
 b E ¼ 1= 2b ð14Þ
qe ¼ x m e 2 ð11Þ
where xm (mg g− 1) is the maximum adsorption capacity, β is the The value of E were found to be 9.1 kJ mol− 1 for Cu(II),
activity coefficient related to mean sorption energy, and ɛ is the 8.8 kJ mol− 1 for Zn(II) and 8.4 kJ mol− 1 for Cd(II) onto TA-
Polanyi potential, which is equal to HTC. Thus the value of E was lied with in 8–16 kJ mol− 1 for all
the metal ions studied here, which indicated that ion exchange
e ¼ RT ln ð1 þ 1=CeÞ ð12Þ mechanism was significant (Erdem et al., 2004).
where R is the gas constant (kJ mol− 1 K− 1) and T is the tem-
perature (K). The adsorption potential is independent of the tem- 3.10. Comparison with other adsorbent
perature that varies according to the nature of adsorbent and
adsorbate. The adsorption capacities (Q0) found out in this study for Cu(II)
The adsorption constants of Cu(II), Zn(II) and Cd(II) ions onto (81.47 mg g− 1), Zn(II) (78.91 mg g− 1) and Cd(II) (74.97 mg g− 1)
TA-HTC were calculated according to Langmuir, Freundlich and were higher than the ones reported in the literature. The values of
DR adsorption models using non-linear regression analysis and Q0 for the adsorption of Cu(II) ions were reported to be 45.9, 31.2
the results are listed in Table 2. Based on correlation coefficient and 2.23 mg g− 1 for adsorption onto ethyleneglycol diglycidyl
(r2) values (N 0.99) and chi-square (χ2) values (b2.0%) for all ether cross-linked chitosan (Nagah et al., 2002), methacrylic/
metals, it was found that the adsorption onto TA-HTC can be best acrylamide monomer mixture grafted fiber (Kun et al., 2006) and
described by the Langmuir model. The qe values obtained from tannic acid immobilized activated carbon (Ucer et al., 2006),
Langmuir model are in agreement with experimental ones (Fig. 9).
The values of Q0 calculated by Langmuir model were found to be Table 3
81.47, 78.91 and 74.97 mg g− 1 for Cu(II), Zn(II) and Cd(II) ions, Chemical composition of wastewaters (mg L− 1)
respectively. The experimental values of maximum monolayer Electroplating Industrial estate Fertilizer industry
capacity (Q) obtained by horizontal extrapolation of knee-point on wastewater for wastewater for wastewater for
Cu(II) Zn(II) Cd(II)
the experimental isotherm to the qe axis. The values of Q were
found to be 81.02, 77.62 and 73.01 mg g− 1 for Cu(II), Zn(II) and Cu2+ 43.3 Zn2+ 20.0 Cd2+ 21.7
Ni2+ 11.7 Cu2+ 20.0 Mg2+ 30.3
Cd(II) ions, respectively; match the values of Q0 calculated by
Na+ 314.7 Glucose 1437.5 Na+ 69.9
Langmuir model. According to Q0 values, the adsorption of the K+ 28.8 Urea 107.3 K 71.5
studied metal ions onto TA-HTC follow the sequence Cu(II) N Zn Mg 611.4 FeSO4.7H2O 24.8 SO2−
4 108.8
(II) N Cd(II). This affinity order is in agreement with the rule for the Ca 43.3 KH2PO4 43.8 CO2−
3 33.4
sequence of complex stability from the Irving–Williams concept Cl− 319.9 pH 6.5 Borate 10.7
SO2− 29.3 NO−3 111.2
based on ligand field or simple-crystal field theory (Cotton et al., 4
CO2−3 89.9 NH4 49.7
1976). According to this rule, the stability of complexes metal Hardness CH3COO− 17.1
cations-oxygen donor groups decreases in the series: Cu(II) N Zn as COD 917.5
(II) N Cd(II). Since the adsorption phenomenon depends on the CaCO3 714.4 pH 5.1
charge density of cations, the diameter of hydrated cations is very COD 413.3
pH 5.1
important. The charges of metal cations are the same (+2); there-
222 T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223

respectively. Erdem et al. (2004), Shukla and Pai (2005) and


Chungsying et al. (2006) have reported the Q0 values of 8.65, 8.02
and 49.94 mg g− 1 for the adsorption of Zn(II) ions onto natural
zeolite, modified jute fiber and multiwalled carbon nanotube,
respectively. The values of Q0 for the adsorption of Cd(II) ions onto
cellulose/ chitosan beads (Zhou et al., 2004), modified chitosan
(Justi et al., 2005) and carbon aerogel (Goel et al., 2006) were found
to be 69.71, 38.50 and 15.53 mg g− 1 respectively. In comparison, it
is clear that Q0 values for Cu(II), Zn(II) and Cd(II) ions obtained in
this study were very high.

3.11. Test with industrial wastewaters


Fig. 11. Regeneration studies of TA-HTC after four cycles.

The utilization of the prepared material as an adsorbent was


assessed by its application in treatment of industrial wastewater.
Simulated industrial estate wastewater with the composition the improvement of process economics. Attempts were made to
(Sirianuntapiboon and Hangsrisuwan, 2007) given in Table 3 was desorb metal ions from spent adsorbent with various eluents
used for Zn(II). Electroplating industry wastewater samples for Cu such as HCl, HNO3, NaCl, NaNO3 and NaAc (0.1 M). The
(II) and fertilizer industry wastewater sample for Cd(II) were desorption efficiency of 0.1 M solutions of HCl, HNO3, NaCl,
collected from local industries situated in the industrial belt of NaNO3 and NaAc was found to be 98.0, 90.6, 80.1, 87.3 and
Cochin city (India). These wastewater samples were characterized 73.2% for Cu(II), 98.4, 92.9, 87.0, 89.7 and 70.1% for Zn(II)
by standard methods (APHA, 1992) and their compositions are and 99.2, 94.4, 91,3, 86.2 and 81.5% for Cd(II), respectively.
given in Table 3. Fig. 10 shows the adsorption of metals as a Thus repeated adsorption/ desorption cycles were performed to
function of adsorbent dose. It is apparent that the amount of examine the reusability and metal ions recovery efficiency of
adsorbed metal increases with increasing the TA-HTC dose. It is the adsorbent (Fig. 11) It is observed that the metal ions ad-
readily understood that the number of available adsorption sites sorption capacity of TA-HTC remained almost constant for the
increases by increasing the adsorbent dose and it, therefore, results four cycles, which indicated that there was no irreversible sites
in the increase of the amount of adsorbed metal ions. The results on the surface of the adsorbent.
also reveal that the treatment of metal ions in wastewater samples
is not significantly different from the results predicted based on 4. Conclusions
single solute batch experiments. Almost complete removal
(100%) of metals was achieved from 1 L of each sample of Cu The ability of tannin-immobilized hydrotalcite to adsorb
(II), Zn(II) and Cd(II) ions with 4.0, 2.0 and 3.0 g L− 1 of adsorbent heavy metal ions such as Cu(II), Zn(II) and Cd(II) from aqueous
respectively. Thus the present study demonstrates that TA-HTC effluents was studied , taking into account kinetic and equili-
can be successfully used for the removal of Cu(II), Zn(II) and Cd brium and mass transfer aspects. FTIR, XRD, SEM, TG/DTG,
(II) ions from industrial wastewaters. Surface area analyzer and potentiometric titrations were carried
out to characterize the adsorbent. The initial and equilibrium
3.12. Regeneration studies adsorption depends on pH. Optimum adsorption was reached at
pH 6.0. Adsorption equilibrium was achieved in approximately
Since metal ions adsorption onto TA-HTC is reversible 3 h. Experimental data were evaluated to find out kinetic
process, it is possible for regeneration of the adsorbent to reuse. characteristics of the adsorption process. Intraparticle diffusion
The repeated use of the adsorbent is likely to be a key factor in was found to take part in adsorption process and it could be
accepted as the primary rate-determining step. Isotherm data
were correlated well with the Langmuir type adsorption and
adsorption capacities were found to be 81.47, 78.91 and
74.97 mg g− 1 for Cu(II), Zn(II) and Cd(II) ions respectively.
Attempts for quantitative removal of Cu(II), Zn(II) and Cd(II)
ions from industrial effluent samples by TA-HTC were made
and satisfactory results were obtained. Repeated adsorption–
desorption study showed that TA-HTC can be effectively used
as an adsorbent for the removal and recovery of heavy metal
ions from aqueous effluents.

Acknowledgment

Fig. 10. Effect of adsorbent dose on the removal of Cu(II), Zn(II) and Cd(II) ions The authors thank the Head, Department of Chemistry,
from industry effluent samples by TA-HTC. University of Kerala, for providing the laboratory facilities.
T.S. Anirudhan, P.S. Suchithra / Applied Clay Science 42 (2008) 214–223 223

References Martin, M.J.S., Villa, M.V., Camazano, M.S., 1999. Glyphosphate–hydrotalcite


interaction as influenced by pH. Clays Clay Miner. 47, 777–783.
Anirudhan, T.S., Suchithra, P.S., 2007. Adsorption characteristics of tannin McKay, G., Bino, M.J., 1988. Adsorption of pollutants from wastewater onto
removal from aqueous solutions and coir industry effluents using calcined activated carbon based on external mass transfer and pore diffusion. Water
Res. 22, 279–286.
and uncalcined hydrotalcites. Ind. Eng. Chem. Res. 46, 4606–4613.
APHA , American Public Health Association, 1992. Standard Methods for the Michelson, C.D., Gideon, P.G., Pace, E.G., Kutel, L.H., 1975. Removal of
Examination of Water and Wastewater, 18th Edn. APHA, AWWA and WEF, soluble mercury from wastewater by complexing techniques. Bull. No. 74.
U. S. Department of Industry, Office of Water Resource and Technology,
Washington DC.
Arco, M.D., Gutierrez, S., Martin, C., Rives, V., 2001. FTIR study of Washington, D.C.
isopropanol reactivity on calcined layered double hydroxides. Phys. Chem. Nagah, W.S.W., Endud, C.S., Mayanar, R., 2002. Removal of copper (II) ions
Phys. 3, 119–126. from aqueous solutions onto chitosan and crosslinked chitosan beads. React.
Funct. Polym. 50, 181–190.
Arslanoglu, F.N., Kar, F., Arslan, N., 2005. Adsorption of dark colored compounds
from peach pulp by using granular activated carbon. J. Food Eng. 68, 409–417. Pavlovic, I., Ulibarr, M.A., Hermosin, M.C., Cornejo, J., 1997. Sorption of an
Baraket, M.A., 2005. Adsorption behavior of copper and cyanide ions at TiO2- anionic Surfactant from water by a calcined hydrotalcite-like sorbent. Fresenius
solution interface. J. Colloid Interface Sci. 291, 345–352. Environ. Bull. 6, 266–271.
Pavlovic, I., Barriga, C., Hermosin, M.C., Cornejo, J., Ulibarr, M.A., 2005.
Barriga, C., Gaiten, M., Pavlovic, J., Ulibari, M.A., Hermossin, M.C., Cornejo, J.,
2002. Hydrotalcite as adsorbent for 2,4,6 trinitrophenol: Influence of the layer Adsorption of acidic pesticides 2, 4-D, clopyratid and picloram on calcined
composition and interlayer anions. J. Mater. Chem. 12, 1027–1032. hydrotalcite. Appl. Clay Sci. 30, 125–133.
Pevez, M.R., Pavlovic, I., Barriga, C., Cornejo, J., Hermosin, M.C., Ulibarr, M.A.,
Chatelet, L., Bottero, J.Y., Yvon, J., Bouchelaghem, A., 1996. Competition between
monovalant and divalent anions for calcined and uncalcined hydrotalcites: anion 2006. Uptake of Cu2+, Cd2+ and Pb2+ on Zn–Al layered double hydroxide
exchange and adsorption sites. Colloids Surf., A Physicochem. Eng. Asp. 111, intercalated with edta. Appl. Clay Sci. 32, 245–251.
167–175. Schwarz, J.A., Driscoll, C.T., Bhanot, A.K., 1984. The zero point charge of
silica alumina oxide suspension. J. Colloid Interface Sci. 97, 55–61.
Chungsying, L., Huanksung, C., Chunti, L., 2006. Removal of Zn(II) from
aqueous solutions by purified carbon nanotubes : kinetic and equilibrium Seida, Y., Nakeno, Y., 2002. Removal of phosphate by layered double hydroxide
studies. Ind. Eng. Chem. Res. 45, 2850–2855. containing iron. Water Res. 36, 1306–1312.
Cotton, F.A., Wilkinsons, G., Murillo, C.A., Bochmann, M., 1976. Advanced Shibi, I.G., Anirudhan, T.S., 2005. Adsorption of Co(II) by a carboxylate-
functionalised polyacrylamide grafted lignocellulosics. Chemosphere 58,
Inorganic Chemistry. Chapter, vol. 20. John Wiley and Sons, New York,
pp. 556–558. 1117–1726.
Das, N.C., Bandyopadhyay, M., 1992. Removal of copper (II) using vermiculate. Shukla, S.R., Pai, R.S., 2005. Adsorption of Cu(II), Ni(II) and Zn(II) on
Water Environ. Res. 64, 852–858. modified jute fibers. Bioresor. Technol. 96, 1430–1438.
Sirianuntapiboon, S., Hangsrisuwan, T., 2007. Removal of Zn2+ and Cu2+ by
Erdem, E., Karapinar, N., Donat, R., 2004. The removal of heavy metal cations
by natural zeolites. J. Colloid Interface Sci. 280, 309–314. a sequencing batch reactor (SBR) system. Bioresor. Technol. 98, 808–818.
Gibert, O., Pablo, J.D., Cortina, J.L., Ayora, C., 2005. Sorption studies of Zn(II) and Tarasov, K.A., Ohare, D., Isopov, V.P., 2003. Solid state chelation of metal ions
by ethylene diammine tetraacetate intercalated in a layer double hydroxide.
Cu(II) onto vegetal compost used as reactive mixtures for in situ treatment of
acid mine drainage. Water Res. 39, 2827–2838. Inorg. Chem. 42, 1919–1927.
Goel, J., Kadirvelu, K., Rajagopal, C., Garg, V.L., 2006. Cadmium (II) uptake Ucer, A., Uyanik, A., Aygun, S.F., 2006. Adsorption of Cu(II), Cd(II), Zn(II),
from aqueous solutions by adsorption onto carbon aerogel using a response Mn(II) and Fe(II) ions by tannic acid immobilized activated carbon. Sep.
Purif. Technol. 47, 113–118.
methodological approach. Ind. Eng. Chem. Res. 45, 6531–6537.
Jansson-Charrier, M., Guibal, E., Roussy, J., Delanghe, B., Cloirec, P.L., 1996. Ulibarri, M.A., Pavlovic, I., Barriga, C., Hermosin, M.C., Cornejo, J., 2001.
Vanadium (IV) sorption by chitosan: Kinetics and equilibrium. Water Res. 30, Adsorption of anionic species on hydrotalcite like compounds: effect of
465–475. interlayer anions and crystallinity. Appl. Clay Sci. 18, 17–27.
Urano, K., Tachikawa, H., 1991. Process development for removal and recovery
Justi, K.C., Favere, V.T., Laranjeira, M.C.M., Navas, A., Peralta, R.A., 2005.
Kinetic and equilibrium adsorption of Cu(II), Cd(II) and Ni(II) ions by chitosan of phosphorous from wastewater by a new adsorbent-2. Adsorption rates and
functionalized with 2[−bis-(pyridyl-methyl)amino methyl]-4-methyl-6-formyl break through curves. Ind. Eng. Chem. Res. 30, 1897–1899.
Yang, L., Shahrivari, Z., Liu, P.K.T., Sahini, M., Tsoksis, T.T., 2005. Removal of
phenol. J. Colloid Interface Sci. 291, 369–374.
Kun, R.C., Soykan, C., Sacak, M., 2006. Adsorption of Cu(II), Ni(II) and Co(II) trace levels of arsenic and selenium from aqueous solutions by calcined and
ions from aqueous solutions by methacrylic acid/ acryl amide monomer uncalcined layered double hydroxide (LDH). Ind. Eng. Chem. Res. 44,
mixture grafted polyethylene terephthalelil fiber. Sep. Purif. Technol. 49, 6804–6815.
You, Y.W., Vance, G.F., Zhao, H.T., 2001. Selenium adsorption on Mg–Al and
107–114.
Leyva-Ramos, R., Rengel-Mendez, J.R., Mendozzo-Barron, J., Fuents-Rubio, Zn–Al layered double hydroxides. Appl. Clay Sci. 20, 13–18.
L., Gerrero-Coronado, R.M., 1997. Adsorption of Cd(II) from aqueous Zhou, D., Zhang, L., Zhou, J., Guo, S., 2004. Cellulose/chitin beads for
solutions onto activated carbon. Water Sci. Technol. 35, 205–211. adsorption of heavy metals in aqueous solution. Water Res. 38, 2643–2650.
Manju, G.N., Gigi, M.C., Anirudhan, T.S., 1999. Hydrotalcite as adsorbent for
the removal of chromium(VI) from aqueous media: Equilibrium studies. Ind.
J. Chem. Technol. 6, 134–141.

You might also like