Sunteți pe pagina 1din 14

611

Removal of Cu2+ and Zn2+ Ions from Aqueous Solution Using Sodium Alginate and Attapulgite Composite Hydrogels
Dajian Huang1,2,3, Wenbo Wang1,3 and Aiqin Wang1,3,* (1) Key Laboratory of Attapulgite Science and
Applied Technology of Jiangsu Provincial, Huaiyin Institute of Technology, Huaian, Jiangsu 223003, P.R. China. (2) University of the Chinese Academy of Sciences, Beijing 100049, P.R. China. (3) Center for Eco-material and Green Chemistry, Lanzhou Institute of Chemical Physics, Chinese Academy of Sciences, Lanzhou, 730000 P. R. China. (Received 3 April 2013; revised form accepted 1 July 2013)

ABSTRACT: A series of sodium alginate-g-sodium acrylate/attapulgite (NaAlg-g-PNaA/APT) composite hydrogels were used as adsorbents to adsorb Cu2+ and Zn2+ from aqueous solutions. The effect of common variables including the content of APT, pH value, contact time and initial concentration of metal ions on the adsorption properties of the composite hydrogels for the adsorption of Cu2+ and Zn2+ ions was investigated using a batch adsorption technique. The results show that the composites evaluated have an excellent adsorption capacity and fast adsorption rate for both Cu2+ and Zn2+ ions. Repeated adsorptiondesorption experimental results show that the composite hydrogels have a good reusability and stability, and 50% of the original adsorption capacity can be retained after being reused five times. The pseudo-second-order equation and the Langmuir equation best describe the adsorption kinetics and isotherms, respectively.

1. INTRODUCTION Heavy-metal pollutants have received considerable attention due to the rapid development of various industrial activities and technologies (Fu and Wang 2011). Most of the heavy-metal ions are highly toxic to human beings and other living organisms even at trace levels, and therefore the effective removal of such heavy-metal ions before discharging industrial effluents is a significant and urgent need. Thus far, numerous technologies, including ion exchange (Plazinski and Rudzinski 2009), adsorption (Xiao et al. 2006), solvent extraction (Alkan and Kara 2004) and membrane separation (Denizli et al. 2001; Abaji 2011), have been developed to remove the metal ions from wastewater. Among them, adsorption is considered as an efficient method because of its advantages such as the flexibility in design and operation and producing superior effluent suitable for reuse without other pollutants (Denizli et al. 2001; Park and Kim 2005; Stafiej and Pyrzynska 2007; Shin et al. 2011). Up to now, many low-cost and environmentally friendly adsorbents including zeolite (Lee et al. 2003), clay (Veli and Alyuz 2007; Weng et al. 2007), sawdust (Yu et al. 2000), biomass (Wang et al. 2010; Iskandar et al. 2011) and chitosan (Ngah et al. 2011) have been used to remove the metal ions from wastewater. However, the adsorption capacities and adsorption rate of these adsorbents are limited. Thus, the development of the adsorbents with high adsorption capacities, fast adsorptiondesorption rate and easy separation and regeneration characteristics has attracted a great deal of interest.
*Author to whom all correspondence should be addressed. E-mail: aqwang@licp.cas.cn

612

D. Huang et al./Adsorption Science & Technology Vol. 31 No. 7 2013

Hydrogels with a three-dimensional cross-linked network structure show relatively higher adsorption capacity towards metal ions and dyes from aqueous solutions (Kavakh et al. 2007; Ju et al. 2009) due to the existence of numerous hydrophilic functional groups (i.e. COOH, NH2 and OH) as chelating sites. In addition, hydrogels also have fast adsorption rate towards target metal ions because even a slight swelling of the hydrogel network may reduce the mass-transfer resistance. However, the use of conventional organic hydrogels based on petroleum monomers increasingly highlighted their disadvantages such as low stability, high production cost and nonbiodegradability, limiting their practical usage. Organic/inorganic composite hydrogels composed of polysaccharide, clay mineral and other bioresources may offer a viable alternative to overcome these drawbacks. Zheng et al. used starchg-poly(acrylic acid)/sodium humate composite hydrogels as the adsorbents to remove Cu2+ from aqueous solutions and the maximum adsorption capacity was found to be 2.61 mmol/g even at a sodium humate content of 20 wt% (Zheng et al. 2010a). Among numerous polysaccharides, sodium alginate (NaAlg) is one of the most potential polysaccharides, as it is renewable, abundant, non-toxic, water soluble, biodegradable and biocompatible (Bhat and Aminabhavi 2007). In addition, NaAlg can be easily modified through various chemical or physical methods such as grafting co-polymerization with hydrophilic vinyl monomers, polymer blending and compounding with other functional components (Hua and Wang 2009). Furthermore, NaAlg also has the ability to adsorb the metal ions from wastewater due to the high activity with carbonyl and hydroxyl groups on its chain (Khotimchenko et al. 2008; Chen et al. 2010). However, to the best of our knowledge, little information has been reported with regard to the adsorption of metal ions by NaAlg and clay hydrogels. Attapulgite (APT) is a kind of hydrated octahedral layered magnesium aluminium silicate mineral with exchangeable cations in its framework channels and reactive OH groups on its surface. It is a type of natural rod-like silicate clay, consisting of two double chains of the pyroxene-type (SiO3)2-like amphibole (Si4O11)6 running parallel to the fibre axis (Liu and Wang 2007a). It can participate in polymerization by its active SiOH to derive new hybrid materials (Liu and Wang 2007b). In our previous work, we had prepared a series of novel sodium alginateg-sodium acrylate (NaAlg-g-PNaA)/APT nanocomposites and the results indicated that the swelling capacities and rates of the nanocomposites were clearly enhanced after introducing APT (Yang et al. 2012). In this paper, NaAlg-g-PNaA/APT composite hydrogels were used as adsorbents to adsorb Cu2+ and Zn2+ from aqueous medium. The parameters influencing the adsorption capacity of the composite hydrogels, such as APT content, contact time, pH and initial concentration of metal ions solution, were evaluated. In addition, the adsorption kinetics and isotherms of the composite hydrogels for metal ions were also discussed.

2. EXPERIMENTAL ANALYSIS 2.1. Materials Sodium alginate (NaAlg) was purchased from Shanghai Chemical Reagents (Shanghai, China). Acrylic acid (AA, chemically pure; Shanghai Shenpu Chemical Factory, Shanghai, China) was distilled under reduced pressure before use. APT was purchased from Xuyi Jiuchuan Colloidal (Jiangsu, China) and was milled and passed through a 320-mesh screen

Cu2+ and Zn2+ Removal Using a Hydrogel Composite

613

(46 m) before use. Ammonium persulphate (APS, analytical grade; Xian Chemical Reagent Factory, China) and N,N-methylene-bis-acrylamide (MBA, chemically pure; Shanghai Chemical Reagent) was used as purchased. Copper acetate monohydrate [analytical grade reagent; Cu (CH3COO)2.H2O, molecular weight (MW) = 199.65] was supplied by Shanghai Reagent (Shanghai, China). Zinc acetate dihydrate [analytical grade reagent, Zn (CH3COO)2.2H2O, MW = 219.50] was supplied by Xilong Chemical Industry Incorporated (Guangdong, China). All other reagents used were of analytical grade and all solutions were prepared with distilled water. 2.2. Synthesis of NaAlg-g-PNaA/APT Composite Hydrogels NaAlg-g-PNaA/APT composite hydrogels were prepared according to the method reported in our previous work (Yang et al. 2012). In a four-necked flask equipped with a stirrer, a thermometer, a reflux condenser and a nitrogen line, NaAlg (1.04 g) was dispersed in 30 ml distilled water and the mixture is stirred to form a transparent sticky solution. After being purged with N2 gas for 30 minutes to remove the dissolved oxygen, the solution was heated to 60 C, and 5 ml aqueous solution of initiator (72 mg APS) was added with continuous stirring. The solution mixture is kept at 60 C for 10 minutes to generate radicals. The temperature of the flask was then decreased to 50 C, and a pre-mixed solution containing 7.2 g of AA (neutralized by adding 8.2 ml of 8.5 mol/l NaOH solution), 28.8 mg cross-linker MBA and a calculated amount of APT powder (0.45, 0.95, 2.05, 3.54 or 4.45 g) was added dropwise into the reactor. The solution was heated to 70 C and maintained for 3 hours to complete the reaction under nitrogen atmosphere. After washing with distilled water for several times, the obtained gel products were dried to constant weight at 70 C. Finally, the dried composite hydrogels were ground and sieved through a 200 mesh (75 m). The NaAlg-g-PNaA hydrogel was prepared according to a similar procedure except without APT addition. 2.3. Adsorption Experiments Batch adsorption experiments were carried out by mixing 0.05 g composite hydrogel with 25 ml of the aqueous solution containing Cu2+ or Zn2+ ions with a known concentration and appropriate pH (Cu2+ solution, 5.5; Zn2+ solution, 6.0). The mixture was then shaken on a thermostatic shaker bath (THZ-98A) with 120 rpm at 30 C for a given period. In order to select the optimum adsorbents for subsequent experiments, the adsorption capacity of the composite with various amounts of APT was first investigated at the contact time of 2 hours. The influence of pH on adsorption was studied by determining the adsorption capability of the hydrogel composite for metal ions at various pH values (Cu2+: 2.0, 3.0, 4.0, 5.0, 5.5 and Zn2+: 2.0, 3.0, 4.0, 5.0, 6.0) at the initial concentration of 650 mg/l. To find the maximum adsorption capacity, the adsorption experiment was conducted at different initial concentrations (1901900 mg/l) and appropriate pH (Cu2+: 5.5 and Zn2+: 6.0) for 2 hours. After adsorption, the solutions of metal ions were separated from the adsorbent by centrifuging the mixture at 5000 rpm for 10 minutes. Both the initial and the final concentrations of metal ions in the solution were measured by ethylenediaminetetraacetic acid (EDTA) titrimetric method using EDTA solution as the standard solution and 0.5% xylenol orange solution as the indicator. The adsorption kinetic experiments were carried out at appropriate pH (Cu2+: 5.5 and Zn2+: 6.0) by examining the adsorption capacities for Cu2+ or Zn2+ ions from aqueous solution (650 mg/l) at pre-determined time intervals.

614

D. Huang et al./Adsorption Science & Technology Vol. 31 No. 7 2013

The equilibrium uptake capacity of the composite hydrogel for metal ion was calculated according to the following equation: q = [V (C0 C)]/m (1)

where q is the amount of metal ion adsorbed at time t or at equilibrium (mg/g); C0 is the initial concentration of metal-ion solution (mg/l); C is the liquid-phase metal-ion concentration at time t or at equilibrium (mg/l); m is the mass of the adsorbent used (g) and V is the volume of metal-ion solution used (l). 2.4. Evaluation of Reusability To identify the reuse value of the composite hydrogel, 0.05 g metal ionloaded sample was agitated with 25 ml HCl solution (0.10 mol/l) at 30 C for 1 hour. After centrifugation, the adsorbent was washed three times with distilled water, and the sample was dried in an oven at 70 C for reuse. The consecutive adsorptiondesorption process was performed five times.

3. RESULTS AND DISCUSSION 3.1. Effect of APT Content on Adsorption The content of clay in the hydrogel is an important parameter influencing the removal amount of metal ions significantly. Figure 1 shows the influence of APT content on the adsorption capacities of the composite hydrogels. It can be seen that the adsorption capacities of the composite hydrogels towards Cu2+ or Zn2+ ions reached maximum when the APT content is 5 wt%, which are slightly higher than that of NaAlg-g-PNaA. These results can be attributed
270

Adsorption capacities (mg/g)

240

210

180

Cu2+ adsorbed Zn2+ adsorbed

150 0 5 10 15 20

APT content (wt%) Figure 1. Effect of APT content on the adsorption.

Cu2+ and Zn2+ Removal Using a Hydrogel Composite

615

to the fact that an appropriate addition of APT would improve the surface area and porosity of the composites, resulting in better adsorption capacity (Zheng and Wang 2009). In addition, OH on the surface of APT could react with AA, which could also improve the polymeric network. However, the content of carboxyl groups in the composite hydrogels will decrease when more OH groups on the surface of APT reacts with AA, which are the main functional groups responsible for the adsorption of metal ions. Therefore, it can be observed that any further increase in the content of APT from 10 to 20 wt% resulted in a decrease in the absorption capacity (Figure 1). Based on this discussion, the following discussion will focus on the NaAlg-g-PNaA/APT composites with an APT content of 5 wt%. 3.2. Effect of pH on Adsorption The pH value of the initial solution is one of the dominant parameter controlling the adsorption process. Figure 2 shows the effect of pH on the adsorption of the composite hydrogels for Cu2+ and Zn2+ ions. It can be seen that the adsorption capacities of the composite hydrogels towards Cu2+ or Zn2+ ions sharply increased with increasing solution pH. This tendency can be explained by the following reason: at lower pH values, the COO groups in the composite hydrogel are protonated as the COOH groups with relatively lower complexing capability towards metal ions, which decreases the adsorption amounts of the composites for the metal ions. With increasing the pH, however, more COO groups are formed in the adsorbents, and therefore, the surface of the composite hydrogel will appear negatively charged, which is favourable for the adsorption of cationic metal ion (Lu et al. 2011). Precipitation of Cu2+ and Zn2+ ions becomes possible at high pH. Therefore, in our study, pH 5.5 and 6.0 were selected as the initial pH value of Cu2+ and Zn2+ solutions for the following adsorption experiments, respectively.
280 240

Adsorption capacities (mg/g)

200 160 120 80


Cu2+

40 0 2 3 4

Zn2+

pH Figure 2. Effect of pH on the adsorption.

616

D. Huang et al./Adsorption Science & Technology Vol. 31 No. 7 2013

3.3. Adsorption Kinetics Figure 3 shows the adsorption kinetic curves of Cu2+ and Zn2+ adsorption on the composite hydrogels at the initial concentration of 650 mg/l. It can be seen that the adsorption of Cu2+ and Zn2+ on the adsorbents took place rapidly and the adsorption equilibrium could be reached within 30 minutes. Therefore, 30 minutes was selected as the adsorption equilibrium time for the subsequent adsorption experiments.
280 240

Adsorption capacities (mg/g)

200 160
Cu2+

120 80 40 0 0 10 20 30 40

Zn2+

50

60

Contact time (min) Figure 3. Effect of contact time on adsorption capacity.

To clarify the adsorption mechanism of Cu2+ and Zn2+ on the composite hydrogels, the pseudofirst-order and pseudo-second-order kinetic models were applied to fit the experimental data. The pseudo-first-order kinetic model developed by Lagergren (Lagergren and Svenska 1898) has been widely used by different authors to predict metal adsorption kinetics, the linear form of which is represented in equation (2) ln(q1e qt) = lnq1e K1t/2.303 (2)

Ho and McKays pseudo-second-order kinetics model may be expressed as follows (Ho and McKay 1999): t/qt = 1/(K2q22e) + t/q2e (3)

where qe (mg/g) and qt (mg/g) are the amounts of the metal ion adsorbed at equilibrium and at time t, respectively; k1 (1/minute) is the rate constant of pseudo-first-order adsorption; k2 (g/mg minute) is the rate constant of the pseudo-second-order adsorption. On the basis of the aforementioned two kinetic models, the correlation coefficients (R2) and the rate constants (k1, k2) obtained from the plot of experiment data are listed in Table 1.

Cu2+ and Zn2+ Removal Using a Hydrogel Composite

617

TABLE 1. Kinetic Parameters for Sorption of Cu2+ and Zn2+ onto Hydrogels Pseudo-first-order model Metal ion Cu2+ Zn2+ qe,experimental (mg g1) 260.34 216.67 q1e (mg g1) 217.90 93.72 k1 (minute1) 0.1375 0.1331 R
2

Pseudo-second-order model q2e (mg g1) 280.11 222.72 k2 103 (g mg1 minute1) 0.9760 3.3652 R2 0.9981 1

0.9902 0.9486

As seen from Table 1, the values of correlation coefficient R2 for pseudo-second-order kinetic model are all above 0.99 for Cu2+ and Zn2+, which are higher than the correlation coefficient R2 for pseudo-first-order kinetic model (0.9902 and 0.9486, respectively, for Cu2+ and Zn2+). Moreover, the calculated q2e values from the pseudo-second-order kinetic model are close to the experimental qe,experimental value. Therefore, it can be concluded that the pseudo-second-order adsorption mechanism dominated the adsorption process and adsorption of Cu2+ and Zn2+ onto the adsorbents were probably controlled by the chemical process. 3.4. Adsorption Isotherm In general, the removal of metal ions from aqueous solutions is dependent on their initial concentration. The adsorption isotherms of Cu2+ and Zn2+ onto the composite hydrogels were studied at various initial concentrations of metal ions (Figure 4). It can be observed that the adsorption capacity of the hydrogel for both Cu2+ and Zn2+ increases sharply when the initial concentration is increased from 190 to 650 mg/l, and then tends to an equilibrium value. Langmuir and Freundlich isotherms were used to test the experimental data in this study. The Langmuir model assumed that the adsorbent has a homogeneous structure, and that all the

350 300

Adsorption capacities (mg/g)

250 200 150 100 50 0 0 400 800 1200 1600 2000


Cu2+ Zn2+

C0 (mg/l) Figure 4. Effect of Cu2+ and Zn2+ concentration on adsorption capacity.

618

D. Huang et al./Adsorption Science & Technology Vol. 31 No. 7 2013

adsorption sites on its surface are identical and energetically equivalent. Mathematically, it is expressed as follows (Langmuir 1918): Ceq/qeq = 1/(b qm) + Ceq/qm (4)

where qeq (mg/g) is the amount adsorbed; Ceq is the equilibrium solute concentration (mg/l) in solution; qm is the maximum adsorption capacity corresponding to the complete monolayer coverage and b is the Langmuir adsorption constant (l/mg). The Freundlich model is valid for heterogeneous system characterized by a heterogeneity factor of 1/n, which describes reversible adsorption and is not restricted to the formation of the monolayer. Mathematically, it is expressed as follows (Freundlich 1906): lnqeq = lnKF + lnCeq 1/n (5)

where KF is the Freundlich isotherm constant (l/g) and 1/n is the heterogeneity factor. The Langmuir and Freundlich parameters obtained from the plot of experiment data are presented in Table 2.
TABLE 2. Langmuir and Freundlich Constants and Correlation Coefficients Associated with Adsorption Isotherms of Cu2+ and Zn2+ onto Hydrogels Langmuir equation Metal ion Cu2+ Zn2+ qm (mg g1) 343.64 307.46 B (l/mg) 0.029 0.013 R2 0.9994 0.9995 K 176.80 99.46 Freundlich equation n 10.84 6.49 R2 0.9461 0.9485

As can be seen from Table 2, the Langmuir model (R2Cu = 0.9976 and R2Zn = 0.9925) appears to fit the equilibrium data better than the Freundlich model (R2Cu = 0.9461, R2Zn = 0.9485), which indicates that the adsorption process is in good agreement with the Langmuir isotherm rather than the Freundlich isotherm. These results reveal the monolayer coverage of metal ions on the surface of the composite hydrogels. In addition, it can be seen from Table 3 that the adsorbents used in the present work have higher adsorption capacities (qm1) than the other adsorbents reported. Especially, when compared with other natural polymer and clay hydrogels, the equilibrium adsorption capacity of NaAlg-gPNaA/APT (5 wt%) is quite high. This depicted the feasibility and advantage of using the hydrogel adsorbents based on NaAlg, and this efficiency can be attributed to the high activity of the carbonyl and hydrogen groups on its chain. 3.5. Regeneration The regeneration ability is a crucial factor for a good adsorbent in addition to its fast adsorption rate and high adsorption capacity from the practical point of view. Consecutive

Cu2+ and Zn2+ Removal Using a Hydrogel Composite

619

TABLE 3. Comparison of the Maximal Adsorption Capacities of Cu2+ and Zn2+ on Different Adsorbents qm1 (mg/g) Adsorbents Acid-activated palygorskite Chitosan/poly(acrylic acid) beads Carboxymethyl chitosan Amidoximated polyacrylonitrile/ organobentonite Poly(VI-co-HEA) hydrogels Starch-graft-acrylic acid Starch-g-poly(acrylic acid)/5% sodium humate MC-g-PNaA/5% APT Ch/IA/MAA hydrogel NaAlg-g-PNaA/5% APT Cu
2+

Zn2+

Reference (Chen et al. 2007) (Dai et al. 2010) (Yan et al. 2011) (Anirudhan and Ramachandran 2008) (Tu et al. 2010) (Guclu et al. 2010) (Zheng and Wang 2010b) (Rheinlander et al. 1998) (Milosavljevic et al. 2011) This paper

32.24 121.55 130.23 77.43 81.37 134.08 181.74 253.5 334.1

65.40 178.65 105.5 286.50

HEA = 2-hydroxyethyl acrylate; MC-g-PNaA = methylcellulose-g-poly(sodium acrylate); VI = N-vinylimidazole.

adsorptiondesorption cycles were performed five times to study the reusability of the composite hydrogels with 0.10 mol/l of HCl solution as the desorption reagent. Adsorption efficiency of the regenerated adsorbents is shown in Figure 5. The adsorption capacities of NaAlg-g-PNaA and NaAlg-g-PNaA/APT (5 wt%) decrease evidently during the first recycle, which can be attributed to the fact that most of the COO groups are changed to COOH while using HCl as the desorption reagent. It can also be seen that the regeneration ratio of NaAlg-g-PNaA/APT (5 wt%) is higher than NaAlg-g-PNaA for both Cu2+ and Zn2+, which indicates that the introduction of APT contributes to improve the regeneration capabilities of the hydrogel adsorbent. This is probably due to the formation of a more stable network of NaAlg-g-PNaA/APT (5 wt%) and
(a)
100

(b)
100
NaAlg-g-PNaA/5% APT NaAlg-g-PNaA

Regeneration ratio (%)

Regeneration ratio (%)

90 80 70 60 50 40 1 2

90

NaAlg-g-PNaA/5% APT NaAlg-g-PNaA

80

70

60

50 3 4 5 1 2 3 4 5

Regeneration times

Regeneration times

Figure 5. Regeneration capabilities of the composite hydrogels on (a) Cu2+ and (b) Zn2+.

620

D. Huang et al./Adsorption Science & Technology Vol. 31 No. 7 2013

that the functional groups of NaAlg, PAA and APT could provide more adsorption sites (Wang et al. 2011). In addition, the acid-desorbed process could increase the adsorption capacity of APT (Chen et al. 2007), which is responsible for the improvement in the regeneration capabilities of the hydrogel adsorbent after the introduction of APT. After five cycles, both NaAlg-g-PNaA and NaAlg-g-PNaA/APT (5 wt%) retained about 50% of their original adsorption capacity for metal ions, indicating that the composite hydrogels are good recyclable adsorbents. 3.6. Adsorption Mechanism To study the adsorption mechanism of composite hydrogel for Cu2+ and Zn2+, the infrared spectra of composite hydrogel before and after the adsorption of Cu2+ and Zn2+ were compared and are shown in Figure 6. As can be seen from Figure 6, the band at 1568 cm1 [Figure 6(a)], related to

(a)

(b)
1717 3445 1701 1408 1568

(c)

1551

1558

3500

3000

2500

2000

1500

1000

500

Wave numbers (cm1) Figure 6. FTIR spectra of the composite hydrogels (a) before and (b) after Cu2+ adsorption, and (c) after Zn2+ adsorption.

the symmetric stretching vibrations of C = O in the COO groups, becomes broad and shifts to lower wave numbers [i.e., 1551 cm1, Figure 6(b), and 1558 cm1, Figure 6(c)] for Cu2+ and Zn2+ adsorption, respectively. In addition, the band at 1719 cm1, assigned to the stretching vibration of C = O of COOH groups, almost disappeared after adsorption. All these results indicated that the groups COOH and COO were involved in the adsorption process (Wang et al. 2009; Zheng and Wang 2010b). Chen et al. had reported that the adsorption of APT for Cu2+ is mainly attributed to the cation exchange of Cu2+ with metal ions between layers of APT and the SiOH of APT (Chen et al. 2007). The adsorption of Zn2+ onto APT may be according to the same mechanism. Therefore, the complexation may be the predominant adsorption mechanism for Cu2+ and Zn2+ onto the hydrogel composites, and the COOH and COO groups may be the main groups responsible for adsorption. Based on the aforementioned discussion, a possible scheme of the

Cu2+ and Zn2+ Removal Using a Hydrogel Composite

621

Cation exchange M2+ -COO-COO-

-COO-COO-COO-COOM2+ -COOH

M2+ -COO2+ -COO- M -COOH -COOM2+ -COO-COO-COO-COOH -COO-COO-COO-COO-COO-COOH M2+ -COOM2+ -COOH M2+ -COOH -COO-COO-

-COO-COOM2+

-COOH -COO-COOM2+ -COOM2+ -COO-COO-COONaAlg -COO-COOM2+ -COOM2+ -COO-COOM2+

M2+ -COO-COO-COOM2+ -COOH -COOH

APT M2+ Cu2+ or Zn2+

PAA

Figure 7. Schematic illustration of the adsorption mechanism of composite hydrogel for Cu2+ and Zn2+.

main adsorption mechanism of NaAlg-g-PNaA/APT (5 wt%) composites for Cu2+ and Zn2+ is shown in Figure 7. 4. CONCLUSIONS In this study, the NaAlg-g-PNaA/APT composite hydrogels were prepared and used as adsorbents to remove the Cu2+ and Zn2+ from the aqueous solution. These adsorbents have high adsorption capacity and fast adsorption rate for Cu2+ and Zn2+. The adsorption capacity of the composite hydrogels increased with increasing pH value of solutions. The pseudo-secondorder equation and the Langmuir equation can describe the adsorption kinetics and isotherms very well, respectively. The metal ions loading on the composite hydrogels can be effectively desorbed by adding 0.10 mol/l HCl solution and the recovered composites have relative higher adsorption capacities towards metal ions even after five adsorptiondesorption processes. The complexing interaction between functional groups and metal ions represents the main adsorption mechanism. The developed composite hydrogel based on renewable natural polymer and natural clays show satisfactory adsorption capacity and rate, and can be used as promising and economic adsorbents for the treatment of wastewater containing heavy metals.

622

D. Huang et al./Adsorption Science & Technology Vol. 31 No. 7 2013

ACKNOWLEDGEMENTS The authors jointly thank the 863 Project of the Ministry of Science and Technology, P.R. China (No. 2013AA031403) and the Science and Technology Support Project of Jiangsu Provincial Science and Technology Department (Nos. BE2012113 and BA2011100).

REFERENCES
Abaji, G. (2011) Adsorpt. Sci. Technol. 29, 169. Alkan, M. and Kara, D. (2004) Instrum. Sci. Technol. 32, 291. Anirudhan, T.S. and Ramachandran, M. (2008) Ind. Eng. Chem. Res. 47, 6175. Bhat, S.D. and Aminabhavi, T.M. (2007) Sep. Purif. Rev. 36, 203. Chen, H., Zhao, Y.G. and Wang, A.Q. (2007) J. Hazard. Mater. 149, 346. Chen, J.H., Li, G.P., Liu, Q.L., Ni, J.C., Wu, W.B. and Lin, J.M. (2010) Chem. Eng. J. 165, 465. Dai, J., Yan, H., Yang, H. and Cheng, R.S. (2010) Chem. Eng. J. 165, 240. Denizli, A., Say, R., Patir, S. and Arica, Y. (2001) Sep. Sci. Technol. 36, 2213. Freundlich, H.M.F. (1906) Z. Phys. Chem. 57A, 385. Fu, F.L. and Wang, Q. (2011) J. Environ. Manage. 92, 407. Guclu, G., Al, E., Emik, S., Iyim, T.B., Ozgumus, S. and Ozyurek, M. (2010) Polym. Bull. 65, 333. Hua, S.B. and Wang, A.Q. (2009) Carbohydr. Polym. 75, 79. Ho, Y.S. and McKay, G. (1999) Process Biochem. 34, 451. Iskandar, N.L., Zainudin N.A.I.M. and Tan, S.G. (2011) J. Environ. Sci. 23, 824. Ju, X.J., Zhang, S.B., Zhou, M.Y., Xie, R., Yang, L. and Chu, L.Y. (2009) J. Hazard. Mater. 167, 114. Kavakh, P.A., Yilmaz, Z. and Sen, M. (2007) Sep. Sci. Technol. 42, 1245. Khotimchenko, M., Kovalev, V. and Khotimchenko, Y. (2008) J. Environ. Sci. 20, 827. Lagergren, S. and Svenska, B.K. (1898) K. Sven. Vetenskapsakad. Handl. 24, 1. Langmuir, I. (1918) J. Am. Chem. Soc. 40, 1361. Lee, M.G., Cheon, J.K. and Kam, S.K. (2003) J. Ind. Eng. Chem. 9, 174. Liu, P. and Wang, T.M. (2007a) J. Hazard. Mater. 149, 75. Liu, P. and Wang, T.M. (2007b) Ind. Eng. Chem. Res. 46, 97. Lu, Q.F., Yu, J., Milosavljevic , N.B., Ristic, M.D., Peric-Grujic, A.A., Filipovic, J.M., Strbac, S.B., Rakocevic, Z.L. and Krusic, T.K. (2011) J. Hazard. Mater. 192, 846. Milosavljevic , N.B., Milas inovic , N.Z., Popovic , I.G., Filipovic , J.M. and Kalagasidis Krus ic , M.T. (2011). Polym. Int. 60: 443. Ngah, W.S.W., Teong, L.C. and Hanafiah, M.A.K.M. (2011) Carbohydr. Polym. 83, 1446. Park, I.H. and Kim, K.M. (2005) Sep. Sci. Technol. 40, 2963. Plazinski, W. and Rudzinski, W. (2009) Environ. Sci. Technol. 43, 7465. Rheinlander, T., Klumpp, E. and Schwuger, M.J. (1998) J. Dispersion Sci. Technol. 19, 379. Shin, K.Y., Hong, J.Y. and Jang, J. (2011) J. Hazard. Mater. 190, 36. Stafiej, A. and Pyrzynska, K. (2007) Sep. Purif. Technol. 58, 49. Tu, J., Zhou, J., Wang, C.F., Zhang, Q.A. and Chen, S. (2010) J. Polym. Sci. Polym. Chem. 48, 4005. Veli, S. and Alyuz, B. (2007) J. Hazard. Mater. 149, 226. Wang, L., Zhang, J.P. and Wang, A.Q. (2011) Desalination 266, 33. Wang, X.H., Zheng, Y. and Wang, A.Q. (2009) J. Hazard. Mater. 168, 970. Wang, X.S., Li, Y.A., Huang, L.P. and Chen, J. (2010) Clean Soil Air Water 38, 500. Weng, C.H., Tsai, C.Z., Chu, S.H. and Sharma, Y.C. (2007) Sep. Purif. Technol, 54, 187. Xiao, J.B., Chen, Q.Y., Chen, X.Q., Jiang, X.Y., Yu, H.Z. and Xu, M. (2006) Adsorpt. Sci. Technol. 24, 547. Yan, H., Dai, J., Yang, Z., Yang, H. and Cheng, R.S., (2011) Chem. Eng. J. 174, 586. Yang, H.X., Wang, W.B. and Wang, A.Q. (2012) J. Dispersion Sci. Technol. 33, 1154.

Cu2+ and Zn2+ Removal Using a Hydrogel Composite

623

Yu, B., Zhang, Y., Shukla, A., Shukla, S.S. and Dorris, K.L. (2000) J. Hazard. Mater. 80, 33. Zheng, Y. and Wang, A.Q. (2009) J. Hazard. Mater. 171, 671. Zheng, Y.A., Hua, S.B. and Wang, A.Q. (2010a) Desalination 263, 170. Zheng, Y.A. and Wang, A.Q. (2010b) Chem. Eng. J. 162, 186.

S-ar putea să vă placă și