11
ORIGINAL ARTICLE Adsorption of Cd 2+ and Cu 2+ ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin Shamsuddeen A. Haladu, Othman Charles S. Al-Hamouz, Shaikh A. Ali * Chemistry Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia Received 23 October 2013; accepted 22 April 2015 KEYWORDS Adsorption; Cross-linked resin; Cadmium; Copper; Metal ion removal Abstract A new cross-linked polyzwitterion (CPZ) was synthesized through cyclocopolymerization of 3-[diallyl(carboxymethyl)ammonio]propane-1-sulfonate (92.5 mol%), and a cross-linker 1,1,4,4- tetraallylpiperazine-1,4-diium chloride (7.5 mol%) in the presence of tert-butylhydroperoxide. The cross-linked polyzwitterion/anion (CPZA) was obtained by the basification of CPZ with NaOH. The adsorption data fit both Temkin and Freundlich isotherm models. The adsorption trend on CPZA is as Cu 2+ > Cd 2+ and both followed pseudo-second-order kinetic model. The negative DG and positive DH values ascertained the spontaneous and endothermic nature of the adsorption process. The effectiveness of a zwitterionic–anionic motif consisting of quaternary nitrogen, sulfonate and carboxylate groups has been tested for the first time for capturing Cd 2+ and Cu 2+ ions at low concentrations. ª 2015 The Authors. Production and hosting by Elsevier B.V. on behalf of King Saud University. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). 1. Introduction Heavy metal ions existence in the environment poses formid- able health problems to various living organisms due to their inherent toxicity which can even lead to death (Srivastava et al., 2011; Waseem et al., 2014). The contamination of water by cadmium occurs majorly via metal plating, cadmium–nickel batteries, mining, smelting and phosphate fertilizers among others. The toxicity of cadmium in human body leads to emphysema, hypertension, renal damage and testicular atro- phy (Rao et al., 2006; Ouadjenia-Marouf et al., 2013) and also enlisted by USEPA as an endocrine disruptor and priority con- trol pollutant (Xin et al., 2007). However, copper pollution emanates from brass manufacture, copper mining, electroplat- ing industries, smelting, Cu-based agri-chemicals (Rao et al., 2006; Ahmad et al., 2012), paint manufacturing and copper polishing. The intake of copper can cause diseases such as vomiting, diarrhea, nausea, liver and kidney damage (Pandey et al., 2009). The threshold level stipulated by WHO for cadmium in drinking water is 0.003 mg/L (WHO, 2008) and that of copper * Corresponding author. Fax: +966 13 860 4277. E-mail address: [email protected] (S.A. Ali). URL: http://faculty.kfupm.edu.sa/CHEM/shaikh/ (S.A. Ali). Peer review under responsibility of King Saud University. Production and hosting by Elsevier Arabian Journal of Chemistry (2015) xxx, xxxxxx King Saud University Arabian Journal of Chemistry www.ksu.edu.sa www.sciencedirect.com http://dx.doi.org/10.1016/j.arabjc.2015.04.028 1878-5352 ª 2015 The Authors. Production and hosting by Elsevier B.V. on behalf of King Saud University. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/). Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd 2+ and Cu 2+ ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin. Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

Adsorption of Cd2+ and Cu2+ ions from aqueous solutions by a cross-linked polysulfonate-carboxylate resin

  • Upload
    iabfu

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Arabian Journal of Chemistry (2015) xxx, xxx–xxx

King Saud University

Arabian Journal of Chemistry

www.ksu.edu.sawww.sciencedirect.com

ORIGINAL ARTICLE

Adsorption of Cd2+ and Cu2+ ions from

aqueous solutions by a cross-linked

polysulfonate–carboxylate resin

* Corresponding author. Fax: +966 13 860 4277.

E-mail address: [email protected] (S.A. Ali).

URL: http://faculty.kfupm.edu.sa/CHEM/shaikh/ (S.A. Ali).

Peer review under responsibility of King Saud University.

Production and hosting by Elsevier

http://dx.doi.org/10.1016/j.arabjc.2015.04.0281878-5352 ª 2015 The Authors. Production and hosting by Elsevier B.V. on behalf of King Saud University.This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+ ions from aqueous solutions by a cross-linked polysulfonate–carboxylaArabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

Shamsuddeen A. Haladu, Othman Charles S. Al-Hamouz, Shaikh A. Ali *

Chemistry Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia

Received 23 October 2013; accepted 22 April 2015

KEYWORDS

Adsorption;

Cross-linked resin;

Cadmium;

Copper;

Metal ion removal

Abstract A new cross-linked polyzwitterion (CPZ) was synthesized through cyclocopolymerization

of 3-[diallyl(carboxymethyl)ammonio]propane-1-sulfonate (92.5 mol%), and a cross-linker 1,1,4,4-

tetraallylpiperazine-1,4-diium chloride (7.5 mol%) in the presence of tert-butylhydroperoxide. The

cross-linked polyzwitterion/anion (CPZA) was obtained by the basification of CPZ with NaOH.

The adsorption data fit both Temkin and Freundlich isotherm models. The adsorption trend on

CPZA is as Cu2+ > Cd2+ and both followed pseudo-second-order kinetic model. The negative

DG and positive DH values ascertained the spontaneous and endothermic nature of the adsorption

process. The effectiveness of a zwitterionic–anionic motif consisting of quaternary nitrogen,

sulfonate and carboxylate groups has been tested for the first time for capturing Cd2+ and Cu2+

ions at low concentrations.ª 2015 The Authors. Production and hosting by Elsevier B.V. on behalf of King Saud University. This is

an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Heavy metal ions existence in the environment poses formid-able health problems to various living organisms due to theirinherent toxicity which can even lead to death (Srivastava

et al., 2011; Waseem et al., 2014). The contamination of water

by cadmium occurs majorly via metal plating, cadmium–nickel

batteries, mining, smelting and phosphate fertilizers amongothers. The toxicity of cadmium in human body leads toemphysema, hypertension, renal damage and testicular atro-phy (Rao et al., 2006; Ouadjenia-Marouf et al., 2013) and also

enlisted by USEPA as an endocrine disruptor and priority con-trol pollutant (Xin et al., 2007). However, copper pollutionemanates from brass manufacture, copper mining, electroplat-

ing industries, smelting, Cu-based agri-chemicals (Rao et al.,2006; Ahmad et al., 2012), paint manufacturing and copperpolishing. The intake of copper can cause diseases such as

vomiting, diarrhea, nausea, liver and kidney damage (Pandeyet al., 2009).

The threshold level stipulated by WHO for cadmium in

drinking water is 0.003 mg/L (WHO, 2008) and that of copper

te resin.

Scheme 1 Synthesis of monomer and cross-linked polymers.

Figure 1 TGA curve of CPZA 6.

Figure 2 IR spectra of (a) cross-linked CPZ 5, (b) CPZA 6, (c)

Cd2+ loaded CPZA 6, (d) Cu2+ loaded CPZA 6.

2 S.A. Haladu et al.

is 0.2 mg/L (WHO, 2003). In view of the above, the treatmentof cadmium and copper contaminated water is of utmost

importance before discharge.Several methods have been employed in the treatment of

metal contaminated water. Among them are Reverse osmosis,

precipitation, dialysis, extraction, adsorption and ionexchange, of which adsorption has received a remarkableattention likely by virtue of the existence of a variety of inex-

pensive green adsorbents (Hong et al., 2007; Shahtaheri

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

et al., 2007). Polyzwitterions bearing pH-responsive groups

have experienced increased attention academically and indus-trially (Dobrynin et al., 2004; Kudaibergenov et al., 2006)and found applications in water treatment, petroleum recov-ery, coatings, drag reduction, cosmetics and viscosification.

Recently, several zwitterionic hybrid materials have been usedeffectively to remove toxic metal ions (Liu et al., 2010; Lianget al., 2006).

In this study, the synthesis of a cross-linked polymer from amonomer having dual functionalities of anionic and zwitteri-onic groups has permitted us to test its efficiency for capturing

Cd2+ and Cu2+ from aqueous solutions at low concentra-tions. This study, to our knowledge, would reveal for the firsttime the employment of a zwitterionic–anionic framework forCd2+ and Cu2+ ions removal.

2. Experimental

2.1. Physical methods

An Elemental Analyzer Series II (Perkin Elmer: Model 2400)

was used to carry out elemental analysis. IR spectra wereobtained using a Perkin Elmer 16F PC FTIR spectrometer.1H and 13C spectra were measured in a JEOL LA 500 MHz

spectrometer with the residual proton resonance of the D2O(at d 4.65 ppm) and dioxane carbon resonance (at d67.4 ppm) as internal standards, respectively. An Atomic

absorption spectroscopy (AAS) model iCE 3000 series(Thermo Scientific) was used to carry out the metal analysis.A TESCAN LYRA 3 (Czech Republic) instrument havingan energy-dispersive X-ray spectroscopy (EDX) detector

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Adsorption of Cd2+ and Cu2+ ions from aqueous solutions 3

(model: X-Max, Oxford) was used to obtain the scanning elec-tron microscopy (SEM) images. A thermal analyzer (STA 429,Netzsch, Germany) was employed to carry out thermogravi-

metric analysis (TGA) using a temperature range20–800 �C (10 �C/min) and air flowing at a rate of 100 mL/min.

2.2. Materials

t-Butylhydroperoxide (TBHP) (70% in water), diallylamineand 1,3-propanesultone from Fluka Chemie AG (Buchs,

Switzerland) were utilized.

2.3. Zwitterionic ester 1 and acid 2

The monomer (1) and the corresponding acid (2) were synthe-sized as reported in Haladu and Ali, (2013).

2.4. 1,1,4,4-tetraallylpiperazine-1,4-diium chloride (4)

Monomer 4, a cross-linker, was prepared as described in Aliet al., (1996).

2.5. Copolymerization of 2 and 4

The monomer 2 (13.9 g, 50 mmol) and cross-linker 4 (1.3 g,4.05 mmol) were dissolved in deionized water followed by

the addition of TBHP (375 mg). The mixture in a closed flask

Figure 3 The adsorption of Cd2+ (1.00 ppm) on CPZA 6 at pH 3 fitte

order kinetic model; (c) intraparticle diffusion model; and (d) adsorpt

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

was stirred for 24 h at 85 �C. After several hours, the magneticstir-bar became static. The gel-like polymer was soaked withwater followed by hot acetone to obtain resin 5 which was

dried in vacuo at 70 �C (8.75 g, 62.2%).

2.6. Transformation of CPZ 5 to cross-linked polyzwitterion/anion (CPZA) 6

CPZ 5 (7.35 g, 26.1 mmol) mixed with sodium hydroxide(1.36 g, 34 mmol) in water (33 ml) was stirred for 24 h. Afteradding excess methanol, resin 6 was filtered and then dried

in vacuo at 65 �C (6.8 g, 87%). At 280 �C, initial decompositionof the resin was shown by TGA.

2.7. Adsorption experiments

CPZA 6 (50 mg) was mixed with aqueous Cd(NO3)2(1 mg Cd L�1) solution (20 mL) at different pH and stirred

for 24 h. The amount of Cd2+ remained in the filtrate wasmeasured by AAS. Using Eq. (1), the adsorption capacity(qCd2þ ) in mg g�1 was calculated as follows:

qCd2þ ¼ðCo � CfÞV

Wð1Þ

where Cf and Co represent the respective final and initial con-centrations (mg L�1) of Cd2+, V and W are the respectivevolume (L) of the solution and the weight (g) of the polymer.

The data from the triplicate experiments varied within 4%.

d to (a) Lagergren first-order kinetic model; (b) Lagergren second-

ion kinetic curve.

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Figure 4 The adsorption of Cu2+ (1.00 ppm) on CPZA 6 at pH 3 fitted to (a) Lagergren first-order kinetic model for; (b) Lagergren

second-order kinetic model; (c) intraparticle diffusion model; and (d) adsorption kinetic curve.

Table 1 Lagergren kinetic model parameters for Cd2+ and Cu2+ ionsa adsorption.

Metal

ion

Temp

(K)

qe, exp(mg g�1)

Lagergren first-order Lagergren second-order Ea

(kJ mol�1)

k1(h�1)

qe, cal(mg g�1)

R2 k2(h�1g mg�1)

hb

(h�1 g�1 mg)

qe, cal(mg g�1)

R2

Cd2+ 294 0.341 4.81 0.016 0.9285 1008 23.4 0.342 1 25.5

308 0.343 6.94 0.014 0.9573 2068 30.4 0.343 0.999

328 0.344 6.52 0.009 0.9813 3009 48.5 0.345 1

Cu2+ 294 0.341 3.66 0.035 0.9860 538 9.43 0.342 1 37.3

308 0.345 4.16 0.009 0.9387 538 13.3 0.345 1

328 0.349 �4.91 0.007 0.9516 1584 16.1 0.345 1

a Initial metal ion concentration 1 mg/L.b Initial adsorption rate h = k2 qe

2.

4 S.A. Haladu et al.

The adsorption capacity (q) at equilibrium and various times isdesignated as qe and qt, respectively, while Ce denotes the equi-librium concentration of the metal ions.

Evaluation of the adsorption isotherm was accomplished in

the Cd2+ concentration range 200 lg L�1 (i.e. ppb) to1000 lg L�1 for 24 h at 294 K. The adsorption kinetic experi-ments were carried out using Cd(NO3)2 solution

(1 mg Cd2+ L�1) at different times at a pH of 3 at 294, 308and 328 K. The data obtained were used to calculate k2 (thepseudo-second-order rate constant) of the adsorption process

and its energy of activation (Ea). The thermodynamic param-eters, DG, DH and DS were determined using the qe and Ce

values. In the same way, the adsorption of Cu2+ ions using

Cu(NO3)2 solution was conducted.

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

2.8. FTIR spectroscopy

FTIR spectra of the loaded and unloaded resins have beenrecorded. About 30 mg of the unloaded resin in 0.1 MCd(NO3)2 solution (20 mL) at pH of 3 was stirred for 4 h, fil-

tered, and dried under reduced pressure to a constant weight.

3. Results and discussion

3.1. Synthesis of CPZA 6

The present resin 6 (Scheme 1) has been prepared bycyclopolymerization reaction. The cyclopolymerization

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Adsorption of Cd2+ and Cu2+ ions from aqueous solutions 5

involving various N,N-diallylammonium monomers has gener-ated various water-soluble cationic polyelectrolytes (Butler,1992). Recently, a resin having aminophosphonate groups

has been synthesized via cyclopolymerization and evaluatedfor its efficacy in the removal of toxic metal ions (AlHamouz and Ali, 2012).

A mixture of monomer 2 (92.5 mol%), and cross-linker 4

(7.5 mol%) in water was cyclocopolymerized using TBHP asan initiator giving CPZ 5 (Scheme 1). Upon treatment of

the CPZ 5 with excess NaOH, CPZA 6 was formed. Amole ratio of �93:7 for monomer 2/cross-linker 4 asdetermined by elemental analysis of CPZ 5 was similar tothe feed ratio.

Three major weight losses were revealed in the TGA curveof CPZA 6 (Fig. 1). The first slow weight loss of 3.0% resultedfrom the release of water trapped in the cross-linked polymer.

The first sharp loss of 22.4% resulted from the release of SO2,the second loss of 28.9% and the third loss of 7.7% were asso-ciated with the respective decarboxylation of the carboxylates

and the degradation of the nitrogenated organic fraction(Martinez-tapia et al., 2000). At 800 �C, the residual masswas found to be 38%.

3.2. Characterization of synthesized materials using FTIR

The IR spectrum of CPZ 5 shows strong absorption bands at1197 and 1041, which are attributed to the sulfonate groups

(Ali et al., 2003). The absorption at 1734 cm�1 indicates thepresence of C‚O stretch of COOH (Al-Muallem et al.,2002b) (Fig. 2a).

The C‚O (of COOH) stretching absorption was missing inthe unloaded resin CPZA 6; however, the absorptions at 1408and 1626 cm�1 were attributed to the symmetric and antisym-

metric stretching of COO� (Fig. 2b). The new band thatappears at 1384 cm�1 (Fig. 2c and d) was assigned to theNO3

� group due to the adsorption process being carried out

using copper and cadmium nitrates (Sahni et al., 1985).Thus, the resin can also act as an anion exchanger due to thepresence of this new band. The increased intensity and broad-ness of the SO3

� and C‚O peaks of the resin loaded with Cu2+

and Cd2+ (Fig. 2c and d) indicated the adsorption of themetal ions (Kolodynska et al., 2009).

Figure 5 The Arrhenius plot for the metal adsorptions.

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

3.3. Adsorption kinetics

The kinetics for the removal of Cu2+ and Cd2+on CPZA 6 ispresented in Figs. 3d and 4d, which depicts the changes inadsorption capacity with adsorption time at different temper-

atures. The adsorption process was relatively fast as it reachedthe equilibrium within 40 min. At higher temperatures, theadsorption capacities increased indicating larger swellingallowing more ions to be diffused and adsorbed on CPZA 6.

3.3.1. Lagergren first-order kinetics

Lagergren first-order kinetics assumes that each metal ionoccupies one adsorption site. The following equation describes

the model:

logðqe � qtÞ ¼ log qe �k1t

2:303ð2Þ

where qt and qe (mg g�1(are the adsorption capacities at time t

(h), and at equilibrium respectively, and k1 is the first-orderrate constant. The intercept and slope of the plots of log(qe–qt) versus t (Table 1, Figs. 3a and 4a) were used to evaluate

qe and k1 experimentally. Only the data of the low time regionwere found to fit the model; hence, the longer time region wasneglected. However, there is a wide disagreement between theequilibrium adsorption capacities using first-order equation

(qe, cal = 0.016 mg g�1 for Cd and 0.035 for Cu at 294 K)and the ones observed experimentally (qe, exp = 0.341 mg g�1

for both Cd and Cu at 294 K). The Lagergren first-order

kinetic equation thus cannot describe the exchange process(Table 1) (Ramesh et al., 2007).

3.3.2. Pseudo-second-order kinetics

The second-order kinetic equation for the adsorption of the metalions can be expressed in linear form by the below equation:

t

qt¼ 1

k2q2eþ t

qeð3Þ

where adsorption capacities are described by qt and qe at therespective time of t and at equilibrium, and k2 represents thesecond-order rate constant. Figs. 3b, 4b and Table 1 proved

that the adsorption of Cd2+ and Cu2+ fits with the second-order kinetic equation implying that the exchange of metalions was likely to be a chemical one (Minceva et al., 2007).

Likewise, the experimentally observed equilibrium adsorptioncapacities were in conformity with those derived from Eq. (3)

3.3.3. Adsorption activation energy

The kinetic rate constants (k2) model was used to calculate the

adsorption energy of activation (Table 1) using the Arrheniusequation (Eq. (4)):

ln k2 ¼ �Ea

2:303RTþ constant ð4Þ

where k2 is the rate constant (g mg�1 h), and R, Ea and Trepresent the gas constant (8.314 J mol�1 K), energy of activa-

tion (kJ mol�1) and temperature (K), respectively. A straightline plot of lnk2 versus 1/T gives a slope of �Ea/R (Fig. 5and Table 1). The energies of activation for the adsorption

of Cd2+ and Cu2+ were found to be 25.5 kJ mol�1 and37.3 kJ mol�1 respectively.

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Table 2 Intraparticle diffusion coefficients and intercept

values for the adsorption of Cd2+ and Cu2+ions at different

temperatures.

Metal

ion

Temp

(K)

kp(mg g�1 h0.5)

Intercept values

(xi)

R2

Cd2+ 294 0.0843 0.301 1

308 0.0436 0.322 1

328 0.0406 0.326 1

Cu2+ 294 0.0843 0.301 1

308 0.0233 0.330 0.9757

328 0.0177 0.336 0.9300

6 S.A. Haladu et al.

3.3.4. Intraparticle diffusion model

The determination of the rate-limiting step paves the way tounderstand the adsorption mechanism. Different adsorptiondiffusion models are constructed based on three steps: first,

film diffusion; second, intraparticle diffusion and third, massaction. According to intraparticle diffusion model, there existinterfacial movements (i.e., film diffusion) of the metal ions

between the adsorbent and the solution, then a subsequentintraparticle diffusion rate-limiting step which delivers the ionsinto the adsorbent pores. The relation of the adsorption capac-

ity and time (Weber and Morris, 1963; Liu et al., 2011) isexpressed as

qt ¼ xi þ kpt0:5 ð5Þ

where kp is the intraparticle diffusion rate constant, qt is the

adsorption capacity at time t and xi is related to the thickness

Figure 6 (a) The pH versus adsorption capacity of CPZA for1.00 pp

of Cd2+ versus adsorption capacity of CPZA 6 at pH 3 for 24 h at 2

Freundlich isotherm for Cd2+ adsorption on CPZA 6.

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

of the boundary layer. In accordance with the Weber–Morris

model (Lin et al., 2011; Weber and Morris, 1963), intraparticlediffusion is the rate-limiting step if a linear fit for the qt versust0.5 plot passes through the origin.

The values of the intercept xi in the initial linear plots forCd2+ were found to be 0.3007, 0.3221, and 0.3255 mg g�1 at21, 35, and 55 �C, respectively (Fig. 3c). The correspondingvalues of 0.3007, 0.3304, and 0.3357 mg g�1 for Cu2+ confirm

the similar trend of increasing xi with the increase in tempera-ture (Fig. 4c) (see Table 2). An intercept exists in the initial lin-ear plot indicating the occurrence of rapid adsorption. The

initial adsorption factor (Ri) was defined in terms of the xi as

Ri ¼ 1� xi

qeð6Þ

where xi and qe are the respective initial adsorption and thefinal adsorption at the longer time. For Cd2+ and Cu2+ ions,the xi and qe values of 0.3007 (Fig. 3c) and 0.341 mg g�1,

respectively, at 21 �C gave an Ri value of 0.12 which meansabout 88% of the initial adsorption has taken place before5 min. The other 12% of adsorption is controlled by intra-

particle diffusion. Hence, the intraparticle diffusion withinthe pores of the resins became the rate-limiting step. Note thatthe horizontal line depicts the equilibrium stage.

3.4. Effect of pH on the adsorption

The pH effect (in the range 3–6) on the adsorption of Cd2+

and Cu2+ on CPZA 6 was studied at 21 �C and a fixed concen-

tration (1 mg L�1) for 24 h. The solution pH was maintained

m Cd2+ and Cu2+ solutions for 24 h; (b) The initial concentration

1 �C; (c) Temkin isotherm for Cd2+ adsorption on CPZA 6; (d)

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Scheme 2 Metal–resin complex.

Adsorption of Cd2+ and Cu2+ ions from aqueous solutions 7

using a buffer of acetic acid/sodium acetate. The adsorption ofmetal ions with the variation of pH is displayed in Fig. 6a. The

best pH was observed to be 3; the adsorption capacity wasfound to increase with further increase in the pH, after theminimum observed at pH 5. Carboxyl and sulfonate groups

may capture Cd2+ and Cu2+ ions by chelation (Scheme 2).The respective pKa values of SO3H and CO2H functionalitiesin CPZA 6 are known to be �2.1 and +2.5 (Al-Muallem

et al., 2002b). As such the repeating unit A is expected to bedissociated to zwitterionic form B even at lower pH values.At pH 3 or even at 2, the acid-base equilibrium, as representedby the equilibrium B � C will involve a certain portion of

CO2� (Zhao et al., 2011) while the SO3H will exist in the con-

jugate base form of SO3�. Lower adsorption of Cd2+ and

Cu2+ at pH < 3 could be attributed to the competition of

H+ ions with Cd2+ and Cu2+ ions for the exchange sites inthe adsorbent to give metal-ion complex D. Above pH 3 theinvolvement of zwitterionic/anionic form D makes the adsorp-

tion faster and more effective. Note that the presentation ofthe metal complex in the cyclic form may not be the preferredchelation mode; the two ligands attached to the divalent metal

ions may as well come from two different repeat units. Atlower pH values species such as monovalent M(Cl)+ may alsoparticipate in the adsorption process (Li et al., 2010; Duonget al., 2006). The metal hydroxides, which are formed owing

to the hydrolysis of the metal ions at higher pH values, cancompete with the resin for the uptake the metal ion(Haghseresht and Lu, 1998). The optimal pH of 3 was chosen

for the adsorption studies.

3.5. Adsorption isotherms

The adsorption capacity of CPZA 6 increases as the initial con-centration of Cu2+ and Cd2+ ions increases, Figs. 6b and 7a.The mechanism of the adsorption was investigated by the anal-ysis of the data using Freundlich, Langmuir and Temkin iso-

therm models. The Langmuir isotherm model(Vijayaraghavan et al., 2006), based on a monolayer adsorp-tion process on homogeneous surfaces where a negligible

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

interaction occurs between the adsorbed molecules, can beexpressed by Eq. (7) as follows:

Ce

qe¼ Ce

Qm

þ 1

Qmbð7Þ

where Ce and qe are the respective metal ion concentrations in

the solution and adsorption capacity of resin, and Qm and bare the Langmuir constants. The constants Qm and b areobtained from the slope and intercept of Ce/qe versus Ce plot.

The Langmuir isotherm model plots for the adsorption ofCu2+ and Cd2+ on the resin have not been presented due tothe non-fitting of the data.

However, Freundlich isotherm model (as expressed by Eqs.

(9) and (10) assumes that a heterogeneous surface with uni-form energy is involved with a multilayer non-idealadsorption:

qe ¼ kfC1=ne ð8Þ

log qe ¼ log kf þ1

nlogCe ð9Þ

where qe and Ce represent the respective equilibrium absorp-

tion capacity of the resin and concentration of metal ion inthe liquid phase, and kf and n denote the Freundlich constants(Table 3), obtainable from the slope and intercept of the log qeversus log Ce in Figs. 6d and 7c. An adsorption is considered

favorable if the values fall within the range of 1 to 10 (Rao andBhole, 2001). The slope (1/n) range of 0–1 is described as a sur-face heterogeneity or a measure of adsorption intensity, which

becomes more heterogeneous when its value approaches zero.A chemisorption process occurs for 1/n value below unity,while 1/n above one implies cooperative adsorption (Table 3)

(Haghseresht and Lu, 1998).The Temkin isotherm equation assumes a linear decrease in

the heat of adsorption of all the molecules in a layer with the

increase in the surface coverage owing to the adsorbent–adsor-bate interactions. The Temkin isotherm (Liu et al., 2011) canbe expressed as Eqs (10)–(12) as follows:

qe ¼RT

blnðaCeÞ ð10Þ

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Figure 7 (a) The initial concentration of Cu2+ versus adsorption

capacity of CPZA 6 at pH 3 for 24 h at 21 �C; (b) Temkin isotherm

for Cu2+ adsorption on CPZA 6; (c) Freundlich isotherm for

Cu2+ adsorption on CPZA 6.

Table 3 Freundlich and Temkin isotherm model constants for

Cd2+ and Cu2+ ions adsorption.

Metal ion kf (mg1�1/n g�1 L1/n) n R2

Freundlich isotherm model

Cd2+ 1.13 1.01 0.9999

Cu2+ 1.35 0.907 0.997

A (L g�1) B (J/mol) R2

Temkin isotherm model

Cd2+ 30.1 0.212 0.9991

Cu2+ 26.9 0.236 0.9814

Figure 8 Vant-Hoff plot.

Table 4 Thermodynamic Data for Cd2+ and Cu2+

adsorption.

Metal

ion

Temperature DG DH DS R2

(K) (kJ/mol) (kJ/mol) (J/mol K)

Cd2+ 294 �1.94 1.36 11.2 0.9563

308 �2.1328 �2.32

Cu2+ 294 �2.08 2.61 16 0.9155

308 �2.31328 �2.63

8 S.A. Haladu et al.

qe ¼RT

blnAþ RT

blnCe ð11Þ

qe ¼ B lnAþ B lnCe ð12Þ

where R, T, and A represent gas constant (8.314 J mol�1 K�1),temperature (K), and equilibrium binding constant (L/g) corre-

sponding to the maximum binding energy, respectively.Constant B (i.e. RT/b) is related to the heat of adsorption.Temkin isotherm constants A and B, calculated from the plot

of qe versus lnCe (Figs. 6c and 7b), are given in Table 3. Theadsorption of Cd2+ and Cu2+ ions fitted well with the Temkinisotherm model as shown in Figs. 6c and 7b. This ascertainsthat the adsorption encounters a heterogeneous surface.

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

3.6. Adsorption thermodynamics

The thermodynamic parameters of the adsorption were alsodetermined. The endothermic nature of the adsorption wasascertained by the increase in the adsorption capacity as

the temperature increases. A log (qe/Ce) versus 1/T plot isshown in Fig. 8. Vant-Hoff equation (Eq. (12)) was employedto calculate the thermodynamic parameters DS, DH and DGthat are tabulated in Table 4 (Ramesh et al., 2007). The neg-

ative values of DG confirmed the spontaneity of the exchangeprocess.

logqeCe

� �¼ � DH

2:303RTþ DS2:303R

ð12Þ

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Adsorption of Cd2+ and Cu2+ ions from aqueous solutions 9

The DG values become more negative with the increase in tem-

peratures, thereby indicating more favorable adsorption pro-cess at the higher temperatures because of greater swellingand increased diffusion of the metal ions into the resin. The

positive values of DH are an indication of endothermic adsorp-tion process. The DS values for the adsorption process werepositive (Table 4) owing to the release of water molecules fromthe metal ions’ large hydration shells.

3.7. SEM and EDX images for CPZA 6 unloaded and loaded

with copper and cadmium ions

SEM has been used to examine both the unloaded and loadedresins. For this purpose, the unloaded resins were stirred in0.1 M Cu(NO3)2 and 0.1 M Cd(NO3)2 at a pH of 3 for 24 h.

The resins were filtered, and dried in vacuo to constant weights.Both resins were then sputter-coated for 4 min with a thin filmof carbon.

The SEM image of the unloaded CPZ 5 and CPZA 6

(Fig. 9a and b) reveals that the surface morphology of CPZ5 is tighter due to the H-bonding among the COOH groups,while the open morphology of CPZA 6 confirms the presence

of Na+ due to the repulsion between the negative charges ofthe CO2

�. The corresponding EDX analysis shows that thecomposition was similar to the proposed one in Scheme 1.

Figure 9 SEM and EDX images for (a) Unl

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

Due to the change in morphology of the SEM images(Figs. 9 and 10) from cracked to smooth, it indicates that theadsorption of cadmium and copper ions had occurred on the

resin. Likewise, the displacement of the sodium ions inCPZA 6 by cadmium and copper ions, Fig 10a and b con-firmed the adsorption of the metal ions.

3.8. Desorption experiment

Efficient use of any adsorbent demands recycling and reuse of

the adsorbent. For this purpose, an experiment was performedas described under Section 2.7 by mixing CPZA 6 (50 mg) withaqueous Cd(NO3)2 (or Cu(NO3)2) (1 mg metal ion L�1) solu-

tion (20 mL) for 24 h. The loaded resin was then recoveredby centrifuging, soaked briefly with water (5 mL) and cen-trifuged again. The recovered resin was dried and treated with0.1 M HNO3 (20 mL) for 24 h for the desorption experiment.

The amount of Cd2+ (or Cu2+) ions desorbed in the filtratewas determined; the efficiency of the desorption process wascalculated by the ratio of desorbed amount of Cd2+ (or

Cu2+) ions to the amount of adsorbed Cd2+ or Cu2+ ions(i.e., qCd2þ or qCu2þ ) in the resin.

The efficiencies of the desorption process, determined to be87% and 89% for Cd2+ and Cu2+ ions, respectively, certifythat the novel cross-linked polymer used in this work is a

oaded CPZ 5 and (b) Unloaded CPZA 6.

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Figure 10 SEM and EDX images for (a) Cd2+ loaded CPZA 6 and (b) Cu2+ loaded CPZA 6.

Table 6 Comparison of the adsorption capacity of the resin

and the adsorption capacity of various adsorbents.

Materials Adsorbents

Capacity (mg/g)

Ref.

Amino acid containing

resin

0.443 for Cu Shaikh et al.

(2013)

Biofilm/GAC system 0.26 for Cd Dianati-Tilaki

and Ali (2003)

Recycled tire rubber 0.0951 for Cu Calisir et al.

(2009)

Chelating polymers with

salicylaldehyde units

0.25 for Cd Amoyaw et al.

(2009)

Present resin 0.345 for Cu,

0.343 for Cd

This work

10 S.A. Haladu et al.

promising candidate as an adsorbent for separation and recov-

ery of metal ions from an aqueous solution.

4. Conclusion

Cyclopolymerization technique provided entry into a novelCPZA which was used to examine the efficiency of a zwitteri-onic/anionic motif in capturing Cd2+ and Cu2+ ions in low

concentrations. At an initial concentration of 1 ppm, therespective removal of Cd2+ and Cu2+ was found to be84.5% and 85.3% (see Table 5). The spontaneity and the

endothermic nature of the adsorption process were ensuredby the negative DGs and positive DHs. A comparison withdifferent types of sorbents in recent references reveals

the excellent metal removal capacity of the current resin(see Table 6).

Table 5 Percent removal at different initial concentrations.

Co (mg L�1) % Removal Cd2+ % Removal Cu2+

0.200 78.1 77.1

0.400 84.6 83.2

0.600 84.5 84.2

0.800 84.7 85.0

1.000 84.5 85.3

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

Acknowledgments

This project was funded by the National Plan for Science,Technology and Innovation (MAARIFAH) – KingAbdulaziz City for Science and Technology – through the

Science & Technology Unit at King Fahd University ofPetroleum & Minerals (KFUPM) – the Kingdom of SaudiArabia, award number (11-ADV2132-04). The authors grate-

fully acknowledge the facilities provided by KFUPM.

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.

Adsorption of Cd2+ and Cu2+ ions from aqueous solutions 11

References

Ahmad, R., Kumar, R., Haseeb, S., 2012. Adsorption of Cu2+ from

aqueous solution onto iron oxide coated eggshell powder:

Evaluation of equilibrium, isotherms, kinetics, and regeneration

capacity. Arab. J. Chem. 5, 353–359.

Ali, S.A., Ahmed, S.Z., Hamad, E.Z., 1996. Cyclopolymerization

studies of diallyl- and tetraallylpiperazinium salts. J. Appl. Polym.

Sci. 61, 1077–1085.

Ali, S.A., Al-Muallem, H.A., Mazumdar, M.A.J., 2003. Synthesis and

solution properties of a new sulfobetaine/sulfur dioxide copolymer

and its use in aqueous two-phase polymer systems. Polymer 44,

1671–1679.

Al Hamouz, O.C.S., Ali, S.A., 2012. Removal of heavy metal ions

using a novel cross-linked polyphosphonate. Sep. Purif. Technol.

98, 94–101.

Al-Muallem, H.A., Wazeer, M.I.M., Ali, S.A., 2002. Synthesis and

solution properties of a new pH-responsive polymer containing

amino acid residues. Polymer 43, 4285–4295.

Amoyaw, P.A., Williams, M., Bu, X.R., 2009. The fast removal of low

concentration of cadmium(II) from aqueous media by chelating

polymers with salicylaldehyde units. J. Hazard. Mater. 170,

22–26.

Butler, G.B., 1992. Cyclopolymerization and Cyclocopolymerization.

Marcel Dekker, New York.

Calisir, F., Roman, F.R., Alamo, L., Perales, O., Arocha, M.A.,

Akman, S., 2009. Removal of Cu(II) from aqueous solutions by

recycled tire rubber. Desalination 249, 515–518.

Dianati-Tilaki, Ali, R., 2003. Study on removal of cadmium from

water environment by adsorption on Gac, Bac and biofilter.

Diffuse Pollution Conference, Dublin.

Dobrynin, A.V., Colby, R.H., Rubinstein, M., 2004. Polyampholytes.

J. Polym. Sci. Part B: Polym. Phys. 42, 3513–3538.

Duong, T.D., Nguyen, K.L., Hoang, M., 2006. Competitive sorption

of Na+ and Ca2+ ions on unbleached kraft fibres––a kinetics and

equilibrium study. J. Colloid Interface Sci. 301, 446–451.

Haghseresht, F., Lu, G., 1998. Adsorption characteristics of phenolic

compounds onto coal-reject-derived adsorbents. Energy Fuel 12,

1100–1107.

Haladu, S.A., Ali, S.A., 2013. Cyclopolymerization protocol for the

synthesis of a new poly(electrolyte-zwitterion) containing quater-

nary nitrogen, carboxylate, and sulfonate functionalities. Eur.

Polym. J. 49, 1591–1600.

Hong, W.U., Yan, J.I.N., Mingbiao, L.U.O., Shuping, B.I., 2007. A

simple and sensitive flow-injection on-line preconcentration cou-

pled with hydride generation atomic fluorescence spectrometry for

the determination of ultra-trace lead in water, wine, and rice. Anal.

Chem. 23, 1109–1112.

Kolodynska, D., Hubicki, Z., Pasieczna-Patkowska, S., 2009. FT-IR/

PAS studies of Cu(II)–EDTA complexes sorption on the chelating

ion exchangers. Acta. Physica. Polonica A 116, 340–343.

Kudaibergenov, S.E., Jaeger, W., Laschewsky, A., 2006. Polymeric

betaines: synthesis, characterization, and application. Adv. Polym.

Sci. 201, 157–224.

Li, Q., Liu, H., Liu, T., Guo, M., Qing, B., Ye, X., Wu, Z., 2010.

Strontium and calcium ion adsorption by molecularly imprinted

hybrid gel. Chem. Eng. J. 157, 401–407.

Liang, W.-jyh., Wu, C.-pang., Hsu, C.-yu., Kuo, P.-lin., 2006.

Synthesis, characterization, and proton-conducting properties of

organic – inorganic hybrid membranes based on polysiloxane

zwitterionomer. J. Polym. Sci. A: Polym. Chem. 44, 3444–3453.

Lin, Y., Chen, H., Lin, K., Chen, B., Chiou, C., 2011. Application of

magnetic particles modified with amino groups to adsorb copper

ions in aqueous solution. J. Environ. Sci. 23, 44–50.

Liu, J., Ma, Y., Xu, T., Shao, G., 2010. Preparation of zwitterionic

hybrid polymer and its application for the removal of heavy metal

ions from water. J. Hazard. Mater. 178, 1021–1029.

Please cite this article in press as: Haladu, S.A. et al., Adsorption of Cd2+ and Cu2+

Arabian Journal of Chemistry (2015), http://dx.doi.org/10.1016/j.arabjc.2015.04.028

Liu, J., Song, L., Shao, G., 2011a. Novel zwitterionic inorganic/

organic hybrids: kinetic and equilibrium model studies on Pb2+

removal from aqueous solution. J. Chem. Eng. Data 56, 2119–2127.

Liu, J., Si, J., Zhang, Q., Zheng, J., Han, C., Shao, G., 2011b.

Preparation of negatively charged hybrid adsorbents and their

applications for Pb2+ removal. Ind. Eng. Chem. Res. 50, 8645–8657.

Martinez-tapia, H.S., Cabeza, A., Bruque, H., Pertierra, P., Garcmh,

S., Aranda, M.A.G., 2000. Synthesis and structure of

Na2[(HO3PCH2)3NH]1.5H2O: the first alkaline Triphosphate. J.

Solid State Chem. 151, 122–129.

Minceva, M., Markovska, L., Meshko, V., 2007. Removal of Zn2+,

Cd2+ and Pb2+ from binary aqueous solution by natural zeolite

and granulated activated carbon. Maced. J. Chem. Chem. Eng. 26,

125–134.

Ouadjenia-Marouf, F., Marouf, R., Schott, J., Yahiaoui, A., 2013.

Removal of Cu(II), Cd(II) and Cr(III) ions from aqueous solution

by dam silt. Arab. J. Chem. 6, 401–406.

Pandey, P., Sambi, S.S., Sharma, S.K., Singh, S., 2009. In: Proceedings

of the World Congress on Engineering and Computer Science, vol.

1, October 20–22.

Ramesh, A., Hasegawa, H., Maki, T., Ueda, K., 2007. Adsorption of

inorganic and organic arsenic from aqueous solutions by polymeric

Al/Fe modified montmorillonite. Sep. Purif. Technol. 56, 90–100.

Rao, M., Bhole, A.G., 2001. Chromium removal by adsorption using

fly ash and bagasse. J. IndianWater Works Assoc. XXXIII 1, 97–100.

Rao, M.M., Ramesh, A., Rao, G.P.C., Seshaiah, K., 2006. Removal of

copper and cadmium from the aqueous solutions by activated

carbon derived from Ceiba pentandra hulls. J. Hazard. Mater.

B129, 123–129.

Sahni, S.K., Bennekom, R.V., Reedijk, J., 1985. A spectral study of

transition-metal complexes on a chelating ion-exchange resin

containing aminophosphonic acid groups. Polyhedron 4, 1643–

1658.

Shahtaheri, S.J., Khadem, M., Golbabaei, F., Rahimi-Froushan, A.,

Ganjali, M.R., Norouzi, P., 2007. Solid phase extraction for

evaluation of occupational exposure to Pb (II) using XAD-4

sorbent prior to atomic absorption spectroscopy. Int. J. Occup. Saf.

Ergon. 13, 137–145.

Shaikh, A.A., Othman, C.S.A., Nouri, M.H., 2013. Novel cross-linked

polymers having pH-responsive amino acid residues for the

removal of Cu2+ from aqueous solution at low concentrations. J.

Hazard. Mater. 248–249, 47–58.

Srivastava, V., Weng, C.H., Singh, V.K., Sharma, Y.C., 2011.

Adsorption of nickel ions from aqueous solutions by nano alumina:

kinetic, mass transfer, and equilibrium studies. J. Chem. Eng. Data.

56, 1414–1422.

Vijayaraghavan, K., Padmesh, T.V.N., Palanivelu, K., Velan, M.,

2006. Biosorption of nickel(II) ions onto Sargassum wightii:

application of two-parameter and three parameter isotherm mod-

els. J. Hazard. Mater. B 133, 304–308.

Waseem, S., Din, M.I., Nasir, S., Rasool, A., 2014. Evaluation of

Acacia nilotica as a non-conventional low cost biosorbent for the

elimination of Pb(II) and Cd(II) ions from aqueous solutions.

Arab. J. Chem. 7, 1091–1098.

Weber, W.J., Morris, J.C., 1963. Kinetics of adsorption on carbon

from solutions. J. Sanit. Eng. Div. Am. Soc. Civ. Eng. 89, 31–60.

WHO, 2003. Copper in Drinking-Water, Guidelines for Drinking-

Water Quality. World Health Organization, Geneva.

WHO, 2008. Guidelines for Drinking Water Quality:

Recommendations. World Health Organization, Geneva, third

ed., vol. 1.

Xin, H., Nai-yun, G., Qiao-li, Z., 2007. Thermodynamics and kinetics

of cadmium adsorption onto oxidized granular activated carbon. J.

Environ. Sci. 19, 1287–1292.

Zhao, G., Li, J.R., Xuemei, Chen, C., Wang, X., 2011. Few-layered

graphene oxide nanosheets as superior sorbents for heavy metal ion

pollution management. Environ. Sci. Technol. 45, 10454–10462.

ions from aqueous solutions by a cross-linked polysulfonate–carboxylate resin.