Вы находитесь на странице: 1из 7

Applied Clay Science 43 (2009) 7985

Contents lists available at ScienceDirect

Applied Clay Science


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / c l a y

Arsenate adsorption by Mg/AlNO3 layered double hydroxides with varying the Mg/Al ratio
Shan-Li Wang a, Cheng Hua Liu b, Ming Kuang Wang b,, Ya Hui Chuang b, Po Neng Chiang b
a b

Department of Soil and Environmental Sciences, National Chung Hsing University, Taichung, 402, Taiwan Department of Agricultural Chemistry, National Taiwan University, Taipei, 106, Taiwan

a r t i c l e

i n f o

a b s t r a c t
Arsenate is one of the main contaminants threatening water supplies all over the world. Layered double hydroxides (LDHs) have large surface areas and high anion exchange capacities, so they may be used to develop rapid and cost-effective methods for scavenging As from contaminated water. In order to enhance arsenate adsorption of layered double hydroxides, it is essential to have a fundamental understanding of the interactions between arsenate and LDHs. In this study, the effects of layer charge density and interlayer nitrate orientation of Mg/AlNO3 LDHs on the preferential adsorption of arsenate were investigated. Mg/AlNO3 LDHs with a Mg:Al ratio of 2:1, 3:1 and 4:1 were synthesized, using a constant-pH co-precipitation method. The adsorption maxima of As on 2:1, 3:1 and 4:1 LDHs were 1.56, 1.08 and 0.36 mmol g 1, respectively, following the order of their layer charge densities. In addition, nitrate orientation in the LDHs plays a signicant role in determining arsenate adsorption. The low arsenate adsorption of 4:1 LDH relative to its anion exchange capacity resulted from the low degree of access arsenate has to the interlayer space, so that arsenate adsorption largely occurred on the external surface of the material. The accessibility of the interlayer of 4:1 LDH to arsenate was restricted by the small basal spacing as a result of the parallel orientation of the interlayer nitrate with respect to the hydroxide sheets. Conversely, interlayer nitrate with an orientation perpendicular to the hydroxide sheets in 2:1 and 3:1 LDHs can be exchanged for arsenate more readily. Consequently, these two LDHs exhibit greater arsenate adsorption capacity, because arsenate can be adsorbed on both their external and interlayer surfaces. The arsenate adsorption of the LDHs also was highly dependent upon the type and concentration of competing anions. The negative effect of the competing anions on arsenate adsorption by various LDHs generally followed the order H2PO N HCO N SO2 N F N Cl. Due to the difference in adsorption mechanism, the arsenate adsorption 4 3 4 of 4:1 LDH is subject to the negative impact of competing anions to a greater extent than those of 3:1 and 2:1 LDH. These results demonstrated that the preferential adsorption of arsenate by LDHs is dependent upon the nitrate orientation in these materials, as a consequence of changing layer charge density. The selective retention of arsenate for wastewater treatment can be maximized by controlling the orientation of interlayer nitrate in Mg/AlNO3 LDHs. 2008 Elsevier B.V. All rights reserved.

Article history: Received 4 March 2008 Received in revised form 27 June 2008 Accepted 2 July 2008 Available online 16 July 2008 Keywords: Arsenate Adsorption Layered double hydroxide Layer charge density

1. Introduction Water supplies worldwide are threatened by arsenic from natural and anthropogenic sources (Mandal and Suzuki, 2002). Arsenic can be present in several oxidation states (3, 0, 3 and 5), but only As(V) and As (III) occur prominently in the environment (Smedley and Kinniburgh, 2002). The major sources of As contamination are percolating water from mines, wood preservatives, and agricultural chemicals (Korte and Fernando, 1991; Smedley and Kinniburgh, 2002). In addition to manmade pollutions, elevated levels of As also occur naturally in water and soil in certain areas of the world, as a result of leaching from As-bearing

Corresponding author. Tel.: +886 2 3366 4808 (O); fax: +886 2 2366 0751. E-mail address: mkwang@ntu.edu.tw (M.K. Wang). 0169-1317/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.clay.2008.07.005

minerals (Korte and Fernando, 1991; Smedley and Kinniburgh, 2002). Arsenic is highly carcinogenic after long-term or high-dose exposure, and the current US-EPA maximum contamination level (MCL) of As in drinking water is 10 parts per billion (US-EPA, 2001). Compliance with this strict standard requires technologies that are highly efcient in removing As from contaminated water. Adsorption method is considered one of the most promising technologies for removing As from water, because of its removal effectiveness, low costs and easy equipment handling (Jiang, 2001; Bissen and Rimmel, 2003). Under most environmental conditions, adsorption by iron and aluminum (hydr)oxides is an important mechanism for immobilizing As, because the oxyanions of As (i.e., arsenate and arsenite) exhibit relatively high afnity for these (hydr) oxide minerals (Smith et al., 1998; Smedley and Kinniburgh, 2002). Thus, many investigators have proposed using these (hydr)oxide

80

S.-L. Wang et al. / Applied Clay Science 43 (2009) 7985

minerals to remove As from contaminated water (Edwards, 1994; Raven et al., 1998; Carlson et al., 2002; Garcia-Sanchez et al., 2002; Gimenez et al., 2007). According to Edwards (1994), the adsorption of arsenate and arsenite onto Fe and Al (hydr)oxides ranges from 0.04 to 0.7 moles per mole (hydr)oxide, depending upon the pH of the solutions and the surface area and particle size of the adsorbents. Among (hydr)oxides with varying degrees of crystallinity, poorly crystalline (hydr)oxides usually exhibit higher As adsorption capacity than their well crystalline counterparts, due to their greater surface area and surface reactivity. For example, ferrihydrite (a poorly crystalline Fe hydroxide) exhibited a maximum arsenate adsorption capacity of 1.23.5 mmol g 1 (Raven et al., 1998), compared with the value of 7.6 10 2 mmol g 1 reported for hematite (a well crystalline Fe oxide) (Gimenez et al., 2007). Thus, to enhance the efcacy of As adsorption, using poorly crystalline (hydr)oxides is better than using their well crystalline counterparts. The major concerns associated with using poorly crystalline materials are their relative instability and high solubility after they are applied to treat contaminated water. Secondary contamination of water supplies can result from the dissolution of contaminant-bearing (hydr)oxides. Arsenic also can be mobilized in the presence of phosphate or organic matter, both of which compete for adsorption sites on (hydr)oxides (Welch et al., 2000; Simeoni et al., 2003). Layered double hydroxides (LDHs) have relatively large surface areas and high anion exchange capacities so they have been studied extensively to evaluate their potential as adsorbents of anionic contaminants, like 2,4-dichlorophenoxyacetate (Pavlovic et al., 2005; Chao et al., 2008), benzoate (Cardoso et al., 2003), phosphate (Seida and Nakano, 2002; Das et al., 2006; Wang et al., 2007), chromate (Goswamee et al., 1998; Kovanda et al., 1999; Tzou et al., 2006; Wang et al., 2006), selenate (You et al., 2001a), and arsenate (Kovanda et al., 1999; Dousova et al., 2003; Gillman, 2006; Yang et al., 2006). The major group of LDHs can be depicted by the formula: h M2 M3 OH2 1x x ix Ay dnH2 O x=y

where M2+ and M3+ are divalent and trivalent metal cations (Cavani et al., 1991). The positive layer charges result from the isomorphous substitution of trivalent cations for divalent cations, and the M3+/(M2+ +M3+) molar ratio (i.e., x in the above formula) ranges from 0.18 to 0.33, corresponding to a layer charge density ranging from 2 to 4 (+) nm 2, respectively (Brindley and Kikkawa, 1979). The positive charges of LDHs are balanced by interlayer anions that can be exchanged for other anions. Thus, unlike other anion adsorbents, like Al and Fe (hydr)oxides, on which contaminants are adsorbed only on their external surfaces, the internal surface of each individual hydroxide sheet in LDHs can provide binding sites for contaminants in addition to those on their external surfaces. Because of this, the adsorption capacities of LDHs are comparable with those of poorly crystalline Fe and Al (hydr)oxides, but LDHs do not have the drawback of structural instability due to their better crystallinity. Previous studies have demonstrated that Mg/Al LDHs have a maximum arsenate adsorption of 5615 mg g 1, depending upon the types of cations and anions present in the structures of the LDHs, and upon experimental conditions such as the ionic strength, pH, and initial As concentration of the solutions (Kovanda et al., 1999; Dousova et al., 2003; Gillman, 2006; Yang et al., 2006). Hence, LDHs may be useful for scavenging As in drinking water and industrial wastewater. Despite the impressive performance of LDHs in scavenging As under laboratory conditions, there are some uncertainties with respect to the As adsorption efciency of LDHs under realistic environmental conditions with complex chemical compositions. One concern pertaining to the desirable function of LDHs stems from the competition for the positively-charged binding sites of LDHs between contaminants of concern (i.e., arsenate in this study) and anions, such as Cl, NO and HCO, which are highly concentrated in industrial 3 3

efuents and polluted water bodies. In order to promote highly selective binding of arsenate by LDHs, it is necessary to develop a fundamental understanding of the reaction mechanisms between As and LDHs. This work focused on how the Al3+/Mg2+ molar ratio and nitrate orientation of Mg/AlNO3 LDHs determine their arsenate adsorption efciency and selectivity. A previous study demonstrated that the orientation of the interlayer nitrate in Mg/Al LDHs is dependent on the layer charge density of the materials (Wang and Wang, 2007). The interlayer nitrate of Mg/Al LDHs has an orientation that is parallel to hydroxide layers with a low charge density (x = 0.20), and an orientation that is perpendicular to hydroxide layers with a high charge density (x = 0.33). In Mg/AlNO3 LDH, whose charge density is somewhere between the above two end-member situations, nitrate ions with both parallel and perpendicular orientations coexist in the same structure. The current study was conducted to further investigate the effects of nitrate orientation in Mg/AlNO3 LDHs with varying layer charge densities, in terms of the arsenate adsorption capacities and mechanisms of these materials. Arsenate in aqueous solutions has various protonation states from H3AsO4 to AsO3 4 (pKa = 2.22, 6.98, and 11.53) (Cullen and Reimer, 1989). Because H2AsO is the predominant species of arsenate in the pH range of 4 natural surface environment, it was selected to be the target contaminant to study its reaction with LDHs. Meanwhile, since industrial efuents and polluted water bodies usually contain other anions (such as Cl, HCO and SO2) that can compete with arsenate for 3 4 the surface binding sites of LDHs, the selective binding of arsenate by various LDHs must be investigated to determine whether these materials are suitable for use in real-life situations. Thus, the adsorption performance of various LDHs for arsenate removals in water containing Cl, NO, HCO, H2PO and SO2 was also investi3 3 4 4 gated. The results of adsorption experiments augmented with the structural information provided by X-ray diffraction (XRD) and X-ray absorption spectroscopy (XAS) were used to determine the nature of the interactions between LDHs and arsenate and to provide insights into the preferential binding of As. This information is essential for designing an optimum scavenger for selectively binding arsenate in water that can be extended to other important anionic contaminants in the environment. 2. Materials and methods 2.1. Samples preparation Mg/AlNO3 LDHs were synthesized using a constant-pH coprecipitation method (Cavani et al., 1991). Five hundred mL of deionized water were added to a Teon reaction vessel and the pH of the water was adjusted to 10 using a 2 M NaOH solution. Nitrate salt solutions containing Al(NO3)3 9H2O and Mg(NO3)2 6H2O (J.T. Baker, Inc., NJ.) were prepared with a total cation concentration of 1.0 M, and Mg/Al molar ratios of 2, 3 and 4 (hereafter referred to 2:1, 3:1 and 4:1 LDHs, respectively). Each of the solutions was added drop-wise into the vigorously-stirred de-ionized water in the Teon reaction vessel, while a 2 M NaOH solution was added simultaneously to maintain a constant pH of 10 using a Radiometer Analytical TIM 856 titrator. The synthetic process was carried out at 25 C under N2 gas purging. After the drop-wise addition of nitrate salt solution was complete, the suspensions were aged at 65 C for 18 h. Precipitates were ltered and washed with de-ionized water to remove excessive salts. The collected solids were freeze-dried and stored in glass vials prior to use. 2.2. Characterization of LDHs The Al and Mg contents of the LDHs were determined using a Perkin Elmer Optima 2000DV ICP-OES instrument after dissolving the samples in aqua regia. The carbon, nitrogen and hydrogen contents of the LDHs were obtained in an elemental analyzer (EA, Heraeus

S.-L. Wang et al. / Applied Clay Science 43 (2009) 7985

81

VarioE1-III). The XRD patterns of samples were obtained on a Rigaku GeigerFlex diffractometer with Cu K radiation, using a scanning rate of 2o2 min 1. 2.3. Adsorption isotherms Batch adsorption studies were performed at 25 C to obtain the adsorption isotherms. Arsenate solutions used for the adsorption experiments were prepared from analytical-grade Na2HAsO4 7H2O (J.T. Baker, Inc.). The initial pH of the arsenate solutions was set to 5.0, and 100 mL arsenate solutions of specic concentrations (i.e., 0.1, 0.3, 0.5, 0.75, 1.0, 1.5 and 2.0 mM) were added to polypropylene bottles. Subsequently, 50 mg of LDHs were added to the arsenate solution in each of the polypropylene bottles, and the suspensions were shaken at 200 rpm for 2 h. After the suspensions were passed through 0.22-m membrane lters to collect the ltrates, arsenate concentrations in the ltrates were determined using ICP-OES. All experiments were conducted in triplicate, and data from the triplicate analyses then averaged. The adsorption isotherms were subjected to the Langmuir equation to determine the maximum adsorption of arsenate for each LDH. 2.4. Effects of competing anions Arsenate solutions containing different competing anions (i.e., Cl, F , SO2, HCO, and H2PO) were prepared by dissolving various 4 3 4 amounts of their sodium salts (J.T. Baker, Inc.) in 2 mM arsenate solutions, and the resultant competing anion/arsenate ratios were 0.2, 1, 5, 10, and 50. Fifty mg LDHs then were added to each of the solutions in 100 mL PP bottles and the suspensions shaken at 200 rpm for 2 h. After ltering the suspensions through 0.22 m membrane lters, the concentrations of arsenate in the ltrates were determined by ICP-OES.

by 3, and then Fourier transformed within a selected data range to produce radial distribution functions (RDF) to isolate the distances from As to different atomic shells. The structural parameters of NaAsO2 and Na2HAsO4 were obtained from the inorganic crystal structure database (ICSD) and entered into the FEFF8.20 (Ankudinov et al., 1998) to compute the theoretical amplitude and phase functions for different tting paths. The geometrical parameters of these paths (i.e., coordination number, inter-atomic distance between As and other atoms, DebyeWaller factor) are extracted by tting the experimental data to FEFF calculations. 3. Results and discussion 3.1. Characterization of LDHs The results of chemical analysis for the LDH samples are shown in Table 1. The resultant molar fractions of Al3+ (i.e., xexp) determined for the synthesized 2:1 LDH, 3:1 LDH and 4:1 LDH were 0.33, 0.26 and 0.21, respectively, which were in good agreement with the initial values (xcal) of the starting solutions (Table 1). Because the positive layer charges of the LDHs result from the isomorphous substitution of Al3+ for Mg2+ in the hydroxide layer [AlxMg1 x(OH)2]x+, the x value represents the structural positive charge density of the LDHs. The numbers of counter-balancing nitrate molecules in these samples were stoichiometrically close to the excess positive charges in the hydroxide sheets, and there was a small amount of contaminated CO2 3 detected in the structures. The contribution of the CO2 content in the 3 LDHs to the surface and structural properties of the LDHs was considered negligible. The X-ray diffraction (XRD) patterns of the three LDHs are shown in Fig. 1. The distinct reections observed for the LDHs are (003), (006), (012), (015), (018), (110) and (113), indexed with R3m rhombohedral symmetry (Bellotto et al., 1996). As seen in the results, only one Mg/ AlLDHNO3 phase was present in each sample, and no crystalline Mg (OH)2 or Al(OH)3 phase was detected. As the Al3+/(Al3+ + Mg2+) molar ratio (i.e., the x value) of the LDHs was varied, a signicant change in the values of the corresponding (003) reection was observed (Fig. 1). The d(003) values of 2:1 LDH, 3:1 LDH and 4:1 LDH, corresponding to the basal spacing of two consecutive brucite-like hydroxide layers in the LDHs, were 8.78, 8.13 and 7.95 , respectively, following the order of 2:1 LDH N 3:1 LDH N 4:1 LDH (Table 1). For the increase in the basal spacing of the LDHs with increasing layer charge density, Xu and Zeng (2001) proposed that the increasing basal spacing results from the change in the corresponding conguration of the interlayer nitrate from a single at-lying molecular layer to multiple layers in the LDH interlayers. Recently, Wang and Wang (2007) suggested different conguration arrangements for the interlayer nitrate in the LDHs, using the results of X-ray

2.5. As K-edge XANE and EXAFS analyses To prepare the samples for As K-edge XAS analyses, 0.5 g of the LDH samples were mixed with 250 mL of 1 mM arsenate solution in 0.01 M NaNO3 at pH 5.0. Suspensions were shaken at 200 rpm for 2 h, and then ltered through a 0.22-um membrane lter. The collected solids were washed with 50 mL de-ionized water ve times on the lter to remove excess salts, freeze-dried, and stored prior to XAS analyses. Arsenic K-edge XAS spectra (including the X-ray absorption near edge structure region or XANES, and the extended X-ray absorption ne structure region or EXAFS) were obtained at the Beamline X17C of the National Synchrotron Radiation Research Center (NSRRC) in Hsinchu, Taiwan. Samples were xed onto an aluminum holder and sealed with Kapton tape. Ionization chambers were used to record As K-edge XAS spectra, ranging from 11,667 to 12,867 eV in the transmission mode and at least two scans were collected for each sample. After acquisition, all scans for each sample were added together. NaAsO2 and Na2HAsO4 chemicals (J.T. Baker, Inc.) were used as the model compounds for determining As(III) and As(V) local environment, respectively. To process the XANES spectra, the before-edge region of each spectrum was tted using a straight line, which was subsequently subtracted from the whole spectrum for baseline correction. Then, all the spectra were normalized to remove As concentration effects on the edge step (Sayers and Bunker, 1988). The EXAFS data were processed using Athena (Ravel and Newville, 2005), following the procedures suggested in Sayers and Bunker (1988). After the spectra were baseline corrected and normalized, energies were recalculated into -space with E0 arbitrary chosen at zero of the main edge rst derivative. To better recognize the contributions of different shell atomic structures to the oscillation function, the data were weighted

Table 1 Structure formulae and basal spacings of Mg/Al-NO3 LDHs and Langmuir parameters derived from their arsenate adsorption isotherms Sample xcala xexpb Structural formula Basal Langmuir parameters spacing () KL r2 b (mmol g 1) (L mmol 1) 1.56 1.08 0.36 47.5 25.7 6.55 1.00 0.99 0.90

2:1LDH 0.33 0.33 [Mg2+ Al3+ (OH)2]0.33+ 8.78 0.67 0.33 2 (NO3)0.29(CO3)0.02 3+ 3:1LDH 0.25 0.26 [Mg2+ Al0.26(OH)2]0.26+ 8.13 0.74 2 (NO3)0.24(CO3)0.01 4:1LDH 0.20 0.21 [Mg2+ Al3+ (OH)2]0.21+ 7.95 0.79 0.21 2 (NO3)0.20(CO3)0.01

a xcal is the molar ratio of trivalent cations to total cations in the initial solutions; xcal = [Al3+]/([Mg2+] + [Al3+]). b xexp is the measured molar ratio of trivalent cations to total cations in the samples.

82

S.-L. Wang et al. / Applied Clay Science 43 (2009) 7985

Fig. 1. XRD patterns of (a) 2:1 LDH, (b) 3:1 LDH and (c) 4:1 LDH.

diffraction and polarized attenuated total reection FTIR spectroscopy. Accordingly, the greater basal spacing of 2:1 LDH may result from the perpendicular orientation of interlayer nitrate ions with respect to the hydroxide sheets of the material. On the other hand, the small basal spacing of 4:1 LDH is due to the horizontal orientation of the interlayer nitrate. In the structure of 3:1 LDH, it was suggested that the interlayer nitrate anions of both the vertical and horizontal orientations coexist within the same particle domain. 3.2. Adsorption isotherms With different Mg2+:Al3+ ratios, the LDHs exhibited different arsenate adsorption characteristics (Fig. 2). Generally, the amount of arsenate adsorbed by the LDHs was dependent on the layer charge density (i.e., xexp). The LDHs exhibited a similar amount of adsorbed arsenate at low arsenate concentration, but the difference among the isotherms became more signicant as the arsenate concentration was increased (Fig. 2). For 2:1 and 3:1 LDHs, the adsorbed amounts of arsenate rapidly increased upon increasing arsenate concentration until reaching the plateau and the adsorption isotherms both were of the L-type according to the classication system proposed by Giles and Smith (1974). This type of adsorption isotherm indicates minimum competition from water and background electrolyte (i.e., NO) for the surface binding sites. By tting the adsorption data with 3

the Langmuir equation, the maximum adsorptions (i.e., the b parameter in Table 1) of arsenate were determined to be 1.56 and 1.08 mmol g 1 for 2:1 and 3:1 LDH, respectively. Comparatively, the characteristics of arsenate adsorption on 4:1 LDH was signicantly different from those of the other two. When the initial arsenate concentration was 0.1 mM, the adsorbed arsenate of 4:1 LDH was 0.15 mmol g 1, which was close to those of 2:1 and 3:1 LDH (Fig. 2). However, upon increasing arsenate concentration, the arsenate adsorption for 4:1 LDH displayed no signicant increase (Fig. 2), and the Langmuir maximum adsorption of arsenate was determined to be 0.36 mmol g 1 (Table 1). When arsenate is adsorbed from aqueous solution by LDHs, the counter-balancing nitrate ions in the structures are replaced by arsenate ions. The anion-exchange process occurs primarily in response to Coulombic attractions between the anionic adsorbate and the positively-charged external and interlayer surfaces of the adsorbents. Therefore, the arsenate adsorption of the three LDHs followed the order of the corresponding layer charges (i.e., xexp), 2:1 LDH N 3:1 LDH N 4:1 LDH. Nonetheless, the maximum amounts of arsenate adsorbed by these three LDHs were not directly proportional to their layer charges (Table 1). The maximum amounts of adsorbed arsenate on 2:1, 3:1 and 4:1 LDHs were in a ratio of 1:0.69:0.23, compared with the ratio of the corresponding xexp values of 1:0.79:0.64. The amount of arsenate adsorbed by 4:1 LDH relative to the corresponding layer charge density was signicantly lower than those by 2:1 and 3:1 LDHs, which cannot be explained by the Coulombic interaction between the adsorbate and adsorbent alone. Thus, structural factors other than the layer charge density of the LDHs also may govern their arsenate adsorption, and the mechanisms behind arsenate adsorption by 4:1 LDH may be different from those for 2:1 and 3:1 LDH. To clarify the structural implications of the differences in the arsenate adsorption of various LDHs, the XRD patterns of the LDH samples before and after arsenate adsorption were investigated (Fig. 3). With the Mg2+:Al3+ molar ratios of 2, 3, and 4, the observed basal spacings of the original LDHs were 8.78, 8.13, and 7.95 , respectively. After interacting with arsenate, the basal spacings of 2:1 and 3:1 LDHs shifted to 8.48 and 7.83 , respectively, while no signicant change was observed for 4:1 LDH (Fig. 3). The low arsenate adsorption of 4:1 LDH (Fig. 2), combined with no change in the basal spacing after arsenate adsorption (Fig. 3), reveals that arsenate has relatively little access to the small interlayer space of 4:1 LDH resulting from the horizontal orientation of the interlayer nitrate with respect

Fig. 2. Adsorption isotherms of arsenate on Mg/Al LDHs (solid concentration = 0.5 g L 1, temperature = 25 C).

Fig. 3. XRD patterns of LDHs before and after arsenate adsorption: (a) 2:1LDH, (b) 2:1LDH-As(V), (c) 3:1LDH, (d) 3:1LDH-As(V), (e)4:1LDH, and (f) 4:1LDH-As(V).

S.-L. Wang et al. / Applied Clay Science 43 (2009) 7985

83

to the hydroxide sheets (Wang and Wang, 2007). Therefore, arsenate may be adsorbed only on the external surface of the 4:1 LDH particles in the range of the initial arsenate concentrations used in this study (Fig. 4). Consequently, the maximum arsenate adsorption was reached at a low initial concentration of arsenate, and was much less than those of 2:1 and 3:1 LDH. Conversely, the changes in the basal spacings of 2:1 and 3:1 LDH after arsenate adsorption indicate the partial replacement of interlayer nitrate ions by arsenate ions. Among the LDHs, 2:1 LDH has the largest basal spacing resulting from the perpendicular orientation of the interlayer nitrate with respect to the hydroxide layers in the material (Wang and Wang, 2007). The large basal spacing facilitates the diffusion of nitrate and arsenate molecules into the interlayer space of 2:1 LDH. Thus, in conjunction with adsorption on the external surface, the interlayer nitrate ions could be replaced by arsenate for the positively-charged binding sites in the hydroxide sheets (Fig. 4), which consequently results in greater arsenate adsorption capacity. It was suggested that 3:1 LDH is an intermediate phase, containing nitrate ions with both parallel and perpendicular orientations (Wang and Wang, 2007). Therefore, we hypothesized that the interlayer nitrate ions with a perpendicular orientation contribute to the arsenate adsorption of 3:1 LDH, while those with a parallel orientation are not replaced and remain in the structure (Fig. 4). Therefore, the arsenate adsorption on 2:1 and 3:1 LDH occurred through the same reaction mechanisms, with different surface availabilities for arsenate.

Fig. 5. The As K-edge XANES spectra of (a) As(III) standard, (b) 2:1 LDH-As(V), (c) 3:1 LDH-As(V), (d) 4:1 LDH-As(V), and (e) As(V) standard.

The local structure of adsorbed arsenate on various LDHs was further determined using the results of XANES and EXAFS. The absorption edges of As adsorbed by various LDHs all were located at 11,875 eV (Fig. 5), which was the same as the value determined for Na2HAsO4, but different from that for NaAsO2. This indicates that there was no reduction of As(V) to As(III) after arsenate adsorption on the LDHs, and that the corresponding local structures of adsorbed arsenate on various LDHs were not different according to their XANES spectra. Fig. 6 demonstrates the 3-weighting oscillation function and RDF prole with peak positions (without phase shift correction), derived from the EXAFS of arsenate adsorbed on 2:1 LDH. The peak positions presented in Fig. 6b correspond to the radial distances from As to atoms at different coordination shells. The tted structural parameters, including the coordination number (CN), inter-atomic distance (R), and DebyeWaller factor (2), are shown in Table 2. The rst coordination shell was attributed to the As(V)-O inter-atomic distance of 1.70 0.02 , with a CN of 4. This is consistent with the molecular structure of an arsenate ion. Furthermore, the inter-atomic distance between As and the second-shell atom was determined to be 2.92 0.07 , with a CN of 2.0. These data cannot be tted to Al or Mg

Fig. 4. Schematic representation of the arsenate adsorption mechanisms for 2:1, 3:1 and 4:1 LDHs.

Fig. 6. (A) k3x(k) spectrum and (B) RDF prole derived from the As K-edge EXAFS of arsenate adsorbed on 2:1 LDH. Solid lines and open circles represent tted and experimental data, respectively (without phase shift correction).

84

S.-L. Wang et al. / Applied Clay Science 43 (2009) 7985

as the second shell atoms from As, but a good t could be obtained if this inter-atomic distance of 2.92 is attributed to that between As and O (Table 2). This result reveals that no direct chemical bond is formed between adsorbed arsenate and the Al or Mg sites in the hydroxide sheets of the LDH. Thus, arsenate ions are adsorbed on the external and internal surfaces of the LDHs by predominantly forming outer-sphere complexes, and the O atom of the second shell may be assigned to the surface OH groups of the hydroxide layers. 3.3. Effects of competing anions on arsenate adsorption When competing anions coexisted with arsenate in the solutions, the adsorption of arsenate ion on various LDHs primarily was affected by the type and concentration of competing anions (Fig. 7). Regardless of the type of competing anion, increasing the concentration of a coexisting competing anion resulted in decrease in arsenate adsorption on different LDHs (Fig. 7). This is due to competition between the coexisting anions and arsenate ion for the positively-charged sites on the external and internal surfaces of the LDHs. Furthermore, the type of co-existing anions determines the extent of selectivity of various LDHs toward arsenate. For example, without the presence of competing anions, 2:1 LDH adsorbed 1.47 mmol g 1 arsenate when the initial arsenate concentration was 1.0 mM. The arsenate adsorption of 2:1 LDH decreased to 1.42, 1.39 and 1.30 mmol g 1 as the Cl/arsenate molar ratio was increased to 1, 10 and 50, respectively (Fig. 7a). Since arsenate adsorption exhibited no signicant decrease upon increasing Cl concentration, the effect of co-existing Cl on the arsenate adsorption of 2:1 LDH was relatively small. Comparatively, a more signicant inuence of increasing the concentration of a competing anion on arsenate adsorption was observed for F, SO2, HCO and 4 3 H2PO . With the competing anion/arsenate molar ratio = 1, the 4 arsenate adsorptions of 2:1 LDH were 1.42, 1.39, 1.14, 0.67 and 0.56 mmol g 1 for the systems containing Cl, F, SO2, HCO and 4 3 H2 PO , respectively. Thus, the effect of competing anion on 4 the arsenate adsorption of 2:1 LDH decreased in the order of H2PO N HCO N SO2 N F N Cl. The same order was also determined 4 3 4 for 3:1 and 4:1 LDH. This order of competing anion effect is consistent with those observed for arsenate adsorption by uncalcined and calcined LDH-CO3 (Yang et al., 2005), as well as those observed for selenite, selenate and arsenite adsorption by various LDHs (Yang et al., 2005; You et al., 2001a,b). As shown in several previous studies, LDHs have a greater afnity toward anions with a higher ionic charge density (Miyata, 1983; Dutta and Puri, 1989; Bontchev et al., 2003). Thus, the co-existence of a competing anion with a higher ionic charge density has a greater negative impact upon the adsorption of a target anion (e.g., arsenate) by LDHs (You et al., 2001a,b; Yang et al., 2005). Among the three LDHs, the competing effect of Cl on arsenate adsorption was more signicant as the layer charge of LDH was lowered. For example, with a Cl/arsenate molar ratio of 5, the amounts of arsenate adsorbed by 2:1, 3:1 and 4:1 LDH were 1.40, 0.93 and 0.22 mmol g 1, respectively, corresponding to 95%, 96% and 73% of their counterparts determined without competition from Cl. Similarly, the decrease in the arsenate adsorption of 4:1 LDH was more

Fig. 7. Adsorption of arsenate on (A) 2:1 LDH, (B) 3:1 LDH, and (C) 4:1 LDH samples with different competing anions (arsenate concentration = 1 mM, solid concentration= 0.5 g L 1, initial pH= 5, temperature = 25 C).

signicant than those of 3:1 and 2:1 LDH, when there was competition from the other anions. As discussed above, the arsenate adsorption of 4:1 LDH occurred on the external surface, while arsenate was adsorbed to the binding sites on both the external and internal surfaces of 3:1 and 2:1 LDH. Thus, the arsenate adsorption of 4:1 LDH is subject to the negative impact of competing anions to a greater extent than what were observed for 3:1 and 2:1 LDH. 4. Conclusions Mg/Al-NO3 LDHs with different Mg2+:Al3+ molar ratios exhibited signicant differences in their arsenate adsorption characteristics due to different nitrate orientations in their interlayers. No evidence of specic adsorption between arsenate and LDH surface was observed in the results of EXAFS analysis; thus, the primary driving force for arsenate adsorption on LDHs is Coulombic attraction between the anionic adsorbate and the positively-charged sites on the surface of the adsorbents. In addition, nitrate orientation in the LDHs plays a signicant role in determining the mechanisms of arsenate adsorption. The interlayer nitrate in 4:1 LDH is parallel to the hydroxide sheets, and the resulting small interlayer space restricts arsenates access to interlayer binding sites. Thus, arsenate is predominantly

Table 2 Structure parameters of arsenate adsorbed by 2:1 LDH derived from the EXAFS spectrum First-shell AsO CNa 4
a b c

Second-shell AsO R()b 1.70 0.02 2(2)c 0.0024 CN 2 R() 2.92 0.07 2(2) 0.0051

Coordination number. Inter-atomic distance. DebyeWaller factor (disorder parameter).

S.-L. Wang et al. / Applied Clay Science 43 (2009) 7985

85

adsorbed on the external surface of 4:1 LDH. On the other hand, with 2:1 and 3:1 LDHs, the interlayer nitrates with an orientation perpendicular to the hydroxide sheets can be replaced by arsenate more readily, relative to their counterparts with a horizontal orientation. As a result, arsenate adsorption by 2:1 and 3:1 LDHs occurs on both the external and interlayer surfaces. Thus, compared with 4:1 LDH, 2:1 and 3:1 LDHs exhibit higher arsenate adsorption and greater preferential afnity toward arsenate when competing with other anions. The results of this study have demonstrated that the selective retention of arsenate from water can be maximized by controlling the orientation of interlayer nitrate in Mg/AlNO3 LDHs. Acknowledgments We thank the National Science Council, Taiwan for the nancial supports (under the grant No. of NSC 91-2313-B-002-361, 92-2313-B002-090, 93-2313-B-002-008, and 93-2313-B-005-055). The authors are grateful to Dr. Jyh-Fu Lee for his assistance in XAS measurements. This research was carried out (in part) at the NSRRC in Hsinchu, Taiwan. References
Ankudinov, A.L., Ravel, B., Rehr, J.J., Conradson, D.D., 1998. FEFF8.20. Phys. Rev. B59, 75657576. Bellotto, M., Rebours, B., Clause, O., Lunch, J.L., Bazin, D., lkaim, E.F., 1996. A reexamination of hydrotalcite crystal chemistry. J. Phys. Chem. 100, 85278534. Bissen, M., Rimmel, F.H., 2003. Arsenica review. Part II: oxidation of arsenic and its removal in water treatment. Acta Hydrochim. Hydrobiol. 31, 97107. Bontchev, R.P., Liu, S., Krumhansl, J.L., Voigt, J., Nenoff, T.M., 2003. Synthesis, characterization, and ion exchange properties of hydrotalcite Mg6Al2(OH)16(A)x (A')2 x4H2O (A, A' = Cl, Br, I, and NO, 2N N0) derivatives. Chem. Mater. 15, 3 36693675. Brindley, G.W., Kikkawa, S., 1979. A crystal-chemical study of Mg, Al and Ni, Al hydroxyperchlorates and hydroxy-carbonates. Am. Miner. 64, 836843. Cardoso, L.P., Tronto, J., Crepaldi, E.L., Valim, J.B., 2003. Removal of benzoate anions from aqueous solution using MgAl layered double hydroxides. Mol. Cryst. Liq. Cryst. 390, 4956. Carlson, L., Bigham, J.M., Schwertmann, U., Kyek, A., Wagner, F., 2002. Scavenging of As from acid mine drainage by schwertmannite and ferrihydrite: a comparison with synthetic analogues. Environ. Sci. Technol. 36, 17121719. Cavani, F., Triro, F., Vaccari, A., 1991. Hydrotalcite-type anionic clays: preparation, properties and applications. Catal. Today 11, 173301. Chao, Y.F., Chen, P.C., Wang, S.L., 2008. Adsorption of 2,4-D on Mg/Al-NO3 layered double hydroxides with varying layer charge density. Appl. Clay Sci 40, 193200. doi:10.1016/j.clay.2007.09.003. Cullen, W.R., Reimer, K.J., 1989. Arsenic speciation in the environment. Chem. Rev. 89, 713764. Das, J., Patra, B.S., Baliarsingh, N., Parida, K.M., 2006. Adsorption of phosphate by layered double hydroxides in aqueous solutions. Appl. Clay Sci. 32, 252260. Dousova, B., Machovic, V., Kolousek, D., Kovanda, F., Dornicak, V., 2003. Sorption of As(V) species from aqueous systems. Water Air Soil Pollut. 149, 251267. Dutta, P.K., Puri, M., 1989. Anion exchange in lithium aluminate hydroxides. J. Phys. Chem. 93, 376381. Edwards, M., 1994. Chemistry of arsenic removal during coagulation and FeMn oxidation. J. Water Work. Assoc. 86, 6478. Garcia-Sanchez, A., Alvarez-Ayuso, E., Rodriguez-Martin, F., 2002. Sorption of As(V) by some oxyhydroxides and clay minerals. Application to its immobilization in two polluted mining soils. Clay Miner. 37, 187194. Giles, C.H., Smith, D., 1974. A general treatment and classication of the solute adsorption isotherm I. Theoret. J. Colloid Interface Sci. 47, 755765.

Gillman, G.P., 2006. A simple technology for arsenic removal from drinking water using hydrotalcite. Sci. Total Environ. 366, 926931. Gimenez, J., Martinez, M., de Pablo, J., Rovira, M., Duro, L., 2007. Arsenic sorption onto natural hematite, magnetite and goethite. J. Hazard. Mater. 141, 575580. Goswamee, R.L., Sengupta, P., Bhattacharyya, K.G., Dutta, P.K., 1998. Adsorption of Cr(VI) in layered double hydroxides. Appl. Clay Sci. 13, 2134. Jiang, J.Q., 2001. Removing arsenic from groundwater for the developing worlda review. Water Sci. Technol. 44, 8998. Korte, N.E., Fernando, Q., 1991. A review of arsenic (III) in groundwater. Crit. Rev. Environ. Control 21, 139. Kovanda, F., Kovacsova, E., Kolousek, D., 1999. Removal of anions from solution by calcined hydrotalcite and regeneration of used sorbent in repeated calcination rehydrationanion exchange processes. Collect. Czechoslov. Chem. Commun. 64, 15171528. Mandal, B.K., Suzuki, K.T., 2002. Arsenic round the world: a review. Talanta 58, 201235. Miyata, S., 1983. Anion-exchange properties of hydrotalcite-like compounds. Clays Clay Miner. 31, 305311. Pavlovic, I., Barriga, C., Hermosin, M.C., Cornejo, J., Ulibari, M.A., 2005. Adsorption of acidic pesticides 2,4-D, clopyralid, picloram on calcined hydrotalcite. Appl. Clay Sci. 30, 125133. Ravel, B., Newville, M., 2005. ATHENA, ARTEMIS, HEPHAESTUS: data analysis for X-ray absorption spectroscopy using IFEFFIT. J. Synchrotron Radiat. 12, 537541. Raven, K.P., Jain, A., Loeppert, R.H., 1998. Arsenite and arsenate adsorption on ferrihydrite: kinetics, equilibrium, and adsorption envelopes. Environ. Sci. Technol. 32, 344349. Sayers, D.E., Bunker, B.A., 1988. Data analysis. In: Koningsberger, D.C., Prins, R. (Eds.), Xray Absorption: Principles, Applications, Techniques of EXAFS, SEXAFS, and XANES. Johns Wiley and Sons, New York, pp. 211253. Seida, Y., Nakano, Y., 2002. Removal of phosphate by layered double hydroxides containing iron. Water Res. 36, 13061312. Simeoni, M.A., Batts, B.D., McRae, C., 2003. Effect of groundwater fulvic acid on the adsorption of arsenate by ferrihydrite and gibbsite. Appl. Geochem. 18, 15071515. Smedley, P.L., Kinniburgh, D.G., 2002. A review of the source, behaviour and distribution of arsenic in natural waters. Appl. Geochem. 17, 517568. Smith, E., Naidu, R., Alston, A.M., 1998. Arsenic in the soil environment: a review. Adv. Agron. 64, 149195. Tzou, Y.M., Wang, S.L., Hsu, L.C., Chang, R.R., Lin, C., 2006. Deintercalation of Li/Al LDH and its application to recover adsorbed chromate from used adsorbent. Appl. Clay Sci. 37, 107114. US-EPA, 2001. National primary drinking water regulations: arsenic and clarications to compliance and new source contaminants monitoring regulated entities. Federal Register Vol. 66, No. 14. Wang, S.L., Wang, P.C., 2007. In-situ XRD and ATR-FTIR study on the molecular orientation of interlayer nitrate in Mg/Al-layered double hydroxides in water. Colloid. Surf. A. 292, 131138. Wang, S.L., Hseu, R.J., Chang, R.R., Chiang, P.N., Chen, J.H., Tzou, Y.M., 2006. Adsorption and thermal desorption of Cr(VI) on Li/Al layered double hydroxide. Colloid. Surf. A. 277, 814. Wang, S.L., Cheng, C.Y., Tzou, Y.M., Liaw, R.B., Chang, T.W., Chen, J.H., 2007. Phosphate removal from water using lithium intercalated gibbsite. J. Hazard. Mater. 147, 205212. Welch, A.H., Westjohn, D.B., Helsel, D.R., Wanty, R.B., 2000. Arsenic in ground water of the United States: occurrence and geochemistry. Ground Water 38, 589604. Xu, Z.P., Zeng, H.L., 2001. Abrupt structural transformation in hydrotalcite-like compounds Mg1 xAlx(OH)2(NO3)x nH2O as a continuous function of nitrate anions. J. Phys. Chem. B. 105, 17431749. Yang, L., Shahrivari, Z., Liu, P.K.T., Sahimi, M., Tsotsis, T.T., 2005. Removal of trace levels of arsenic and selenium from aqueous solutions by calcined and uncalcined layered double hydroxides (LDH). Ind. Eng. Chem. Res. 44, 68046815. Yang, L., Dadwhal, M., Shahrivari, Z., Ostwal, M., Liu, P.K.T., Sahimi, M., Tsotsis, T.T., 2006. Adsorption of arsenic on layered double hydroxides: effect of the particle size. Ind. Eng. Chem. Res. 45, 47424751. You, Y., Vance, G.F., Zhao, H., 2001a. Selenium adsorption on MgAl and ZnAl layered double hydroxides. Appl. Clay Sci. 20, 1325. You, Y., Zhao, H., Vance, G.F., 2001b. Removal of arsenite from aqueous solutions by anionic clays. Environ. Sci. Technol. 22, 14471457.

Вам также может понравиться