Zhipan Wena,
Chaomeng Daib,
Yan Zhua and
Yalei Zhang*a
aState Key Laboratory of Pollution Control and Resource Reuse, College of Environmental Science and Engineering, Tongji University, Shanghai 200092, China. E-mail: zhangyalei@tongji.edu.cn; Tel: +86 21 65985811
bCollege of Civil Engineering, Tongji University, Shanghai 200092, China
First published on 2nd December 2014
Magnetic mesoporous iron manganese bimetal oxide (MMIM) with high specific surface area, pore volume and well interconnected mesopores was synthesized via the nanocasting strategy using KIT-6 as a hard template. The morphology and physicochemical properties of the samples were characterized using SEM-EDS, TEM, XRD, VSM, FT-IR and XPS, etc. The obtained MMIM was used as an adsorbent to remove arsenate from aqueous solutions, and presented excellent performances for As(V) removal. The adsorption equilibrium data were well described by the Freundlich model, and solution pH values affected the removal efficiency of arsenate significantly, which was due to the isoelectric point (IEP) of the MMIM. The adsorption kinetics fitted a pseudo-second order model, and intra-particle diffusion was not the only rate-limiting step. In addition, MMIM exhibited a sensitive magnetic response and could be easily separated and recovered from aqueous solutions with an external magnetic field. Based on analysis results, possible mechanisms were discussed based on a combination of the results to different theoretical adsorption models.
Compared with other conventional methods such as coagulation, reverse osmosis and biological treatment,12–14 adsorption has been proven to be one of the most promising technologies for arsenic remediation, with advantages of having high removal efficiency, simplicity for operation, low cost and high recycle rate without harmful by-products. Among potential adsorbents, iron (hydr)oxides and iron-containing substances have been widely focused on for arsenic species removal due to their strong affinity and high selectivity for inorganic arsenic species in the sorption processes.2 However, the serious drawback for those iron (hydr)oxides and iron-containing adsorbents is low adsorption capacity of arsenic, which limited the usage of these adsorbents. Recently, incorporating some other metal elements such as Zr, Ce, Mn and Co into iron oxides to prepare bimetal oxides is gaining considerable attentions,15–17 since these bimetal oxides not only can inherit the advantages of parent oxides but also show obviously synergistic effect. It is found that iron oxides incorporated with Ce or Mn can significantly increase the arsenic adsorption capacity than the individual metal oxides from aqueous solutions.2,18,19 Magnetic bimetal oxide nanoparticles, possessing high specific surface area, excellent adsorption ability and could be easily separated from solution by using an external magnetic field, have appealing properties by combining the virtues of the each component, which have enhanced properties in comparison with their monometallic nanoparticles.17,20 However, the agglomeration of nanoparticles via the inter-particle dipolar force,21 leading to the loss of size effect and the decrease of surface area,22 which limits their practical application.
Since the discovery of ordered mesoporous silica, mesoporous materials have attracted more attention due to their unique properties such as uniform and adjustable mesopores, tunable and open pore architectures, high specific surface area and large pore volume.23,24 Moreover, the stable and interconnected frameworks of their materials could be easily modified and functionalized without any change in mesostructure,25,26 which could further tailor their adsorption performance. Therefore, these ordered mesoporous silicas are considered to be ideal hard template materials to prepare the ordered non-silica mesoporous metal/bimetal oxides. Compared with common bimetal oxides, bimetal oxides with mesoporous structures have unique physicochemical properties, such as intrinsic high specific surface area and large pore volume, regular and tunable pore size distribution, as well as stable and interconnected frameworks. Such features meet the requirement as excellent adsrobents, not only providing large space and interface capable of accommodating capacious guest species, but also enabling the possibility of specific binding, enrichment and separation.
On the basis of the above considerations, in the present study, a novel magnetic mesoporous iron manganese bimetal oxides (MMIM) with well-defined uniform mesopores was synthesized through the hard template approach, which combined the superiority of bimetal oxides and mesoporous materials. The obtained MMIM sample was used as an adsorbent and presented excellent performance for arsenate removal from aqueous solutions due to the high specific surface area and pore volume, and could be easily separated and recovered from the aqueous solutions using an external magnetic field after the arsenic adsorption. Characterization and analysis were performed by SEM-EDS, TEM, VSM, XRD and XPS before and after adsorption. Based on the results of analysis, possible adsorption mechanism of As(V) removal by MMIM from aqueous solutions was proposed.
The magnetic mesoporous iron manganese bimetal oxides was synthesized as follows: 1.0 g Fe(NO3)3·9H2O, 0.246 g FeCl2·4H2O and 0.311 g Mn(NO3)2·4H2O, which were used as iron and manganese precursors, were dissolved in 20 mL ethanol, and 1.0 g KIT-6 was added into this homogenous solution under vigorous stirring. The mixture was stirred for 2 h at room temperature, and then the ethanol was evaporated at 50 °C. These two synthetic steps were carried out in an anaerobic glove box to prevent Fe(II) oxidized to Fe(III) in mixture. The obtained dry powder was heated to 300 °C and calcined at this temperature for 6 h in high purity nitrogen atmosphere. The impregnation procedure was then repeated with half the amount of the metal precursors, and the precursors@silica composites were calcined 450 °C for 6 h in high purity nitrogen atmosphere; the resulting sample was treated with 2 M NaOH solution to remove the silica template, centrifuged, washed several times with ultrapure water and dried at 40 °C in air, and the final obtained powder was referred to as MMIM. For comparison, KIT-6 templated Fe3O4 was also was synthesized through the KIT-6, and the procedures were similar as the MMIM exception of the manganese precursors. The final obtained powder was denoted as templated Fe3O4.
Aqueous solution samples were immediately centrifuged, and supernatant were filtered with a 0.22 μm membrane, the residual arsenic concentrations were measured by Inductively Coupled Plasma-Optical Emission Spectrometry (ICP-OES, Agilent 720 ES, USA).
Samples | BET (m2 g−1) | Pore size (nm) | Pore volume (cm3 g−1) |
---|---|---|---|
KIT-6 template | 579.60 | 6.25 | 0.777 |
MMIM | 143.88 | 5.88/11.52 | 0.647 |
SEM and EDS images of MMIM are shown in Fig. 1a and b, it was clearly observed that this obtained MMIM possesses large clusters nanoparticles. It was reported that with increasing aging temperature of KIT-6, the interconnectivity between these two channel systems increased.29 In the present study, KIT-6 template obtained at a high aging temperature of 80 °C had higher interconnectivity between these two channel systems, and the nanocasted MMIM resulted in a rather dense structure. In addition, the pore system of the parent KIT-6 template was never completely filled in a nanocasting process, the nanocast material had to aggregate during the curing in parts of the parent template pore system.29 These results made the obtained MMIM in the present study seemed to be hundreds of nanometers. The EDS analysis corresponding to the SEM image clearly revealed the existence of iron and manganese elements in the MMIM, and both the weight ratio and atomic ratio of Fe/Mn were approximately 3/1, which agreed well with the initial Fe/Mn molar ratio. This result suggested that iron and manganese were completely loaded on the channels of the KIT-6 template. The structure and morphology was further investigated with TEM. Highly well-ordered mesoporous structures of MMIM can be clearly seen in Fig. 1c. The lattice fringes were clearly visible with a spacing of 0.254 nm from HRTEM (Fig. 1d), which was in good agreement with the spacing of the (311) planes of Fe3O4.30 No distinct lattices of manganese oxides were clearly observed, suggesting manganese oxides was amorphous in MMIM, which was consistent with WXRD analysis that was discussed later.
The specific surface area (SSA) and mesoporosity parameters of MMIM were investigated by nitrogen adsorption/desorption measurement. The specific surface area of MMIM was obtained by the Brunauer–Emmett–Teller (BET) method, the total pore volume was determined using the adsorbed volume at a relative pressure of 0.99, while pore size distributions of MMIM calculated from the adsorption branch of isotherm because the adsorption branch is highly preferred for pore size calculations and is hardly affected by any tensile strength effect phenomenon.31 Fig. 2a shows the isotherms of MMIM exhibited a type-IV with H1 hysteresis loop, typical characteristic of uniform mesoporous metal oxide prepared by the hard template method. This result revealed that the synthesized MMIM had well-defined uniform mesopores, which was in agreement with the results of TEM images. The detailed texture parameters of MMIM are also shown in Table 1. The obtained MMIM possesses a higher BET surface area and pore volume compared with other reported mesoporous oxides,32–35 and this high surface area and pore volume may be the major reasons for its excellent capacity for arsenate removal from aqueous solutions. The Fig. 2a inset shows the bimodal pore size distribution of the MMIM: the small pore size was approximately 5.9 nm, while the large one was approximately 11.5 nm. Earlier studies have also indicated that KIT-6 template is composed of two channel systems that are connected to each other through micropores channels in the silica walls,29,32,36,37 and the prevalence of these channels varies with the hydrothermal synthesis conditions. Nanocasted replica may grow within the pores of hard template in two ways, which may result in the formation of coupled replicas or uncoupled replicas, when metal oxides grow only in one of the channel systems of the KIT-6 template, the nanocast metal oxides show a bimodal pore size distribution.
![]() | ||
Fig. 2 Nitrogen adsorption/desorption isotherm, pore size distribution (in inset) (a) and WXRD pattern (b) of MMIM. |
Fig. 2b shows the WXRD pattern of MMIM. The diffraction peaks for the sample at 2θ values of 18.2°, 30.0°, 35.4°, 43.0°, 53.2°, 57.0° and 62.6°, corresponding to the reflection planes of (111), (220), (311), (400), (422), (511) and (440), respectively, could be indexed to magnetite Fe3O4 (JCPDS card no. 19-0629). The average crystallite size of the MMIM was calculated from the WXRD by using the Scherrer equation, and the value was approximately 10.7 nm. No principal peaks of manganese oxides appeared in the pattern of the MMIM nanoparticles in Fig. 2b, suggesting that the generated manganese oxides in MMIM may be amorphous, as no obvious diffraction peaks appeared in the WXRD pattern. Pure manganese oxide was also synthesized via the KIT-6 template in order to investigate the phase of the manganese species. Namely, the synthetic method was the same as that for MMIM, but the metal precursors was only manganese nitrate. Fig. S3a† shows the WXRD pattern of the pure manganese oxides, the broad peaks of the manganese oxides pattern indicated a poorly crystalline form, which was the major reason that no obvious diffraction peaks appeared in the WXRD pattern of MMIM. The major phase of manganese oxide were Mn3O4 (JCPDS no. 18-0803) and MnO2 (JCPDS no. 44-0141), suggesting that Mn(II) was oxidized to Mn(III) and Mn(IV) during the high purity nitrogen atmosphere calcinations process, the same result was also deduced from XPS, which would be discussed later. However, compared with the WXRD pattern of the pure manganese oxides, the WXRD pattern of the pure iron oxides (templated Fe3O4), which was shown in Fig. S3b,† was not very different from that of MMIM, and the major phase of iron oxides was magnetite Fe3O4 (JCPDS card no. 19-0629).
The magnetic hysteresis curves measured at room temperature of the MMIM is shown in Fig. 3a. Saturation magnetization (Ms) is a key factor for the successful magnetic separation and the saturation magnetization of MMIM is 33.2 emu g−1, which is lower than the reported for bulk magnetite (92 emu g−1), the smaller particle size and the surface related effects such as surface disorder may be the main reasons for this result,38,39 which is consistent with the obtained results of WXRD and TEM. Due to the nonlinear hysteresis loop with nonzero remnant magnetization (Mr) and coercivity (Hc), MMIM shows well pronounced ferromagnetic properties.40 These results show that the MMIM could possess a sensitive response with an external magnetic field, such as a magnet (Fig. 3a inset), thus providing a potential advantage for the separation and recovery of adsorbent from aqueous solutions after the arsenic adsorption.
FT-IR analysis was conducted to reveal the surface nature of MMIM and As-loaded MMIM, which are shown in Fig. 3b. All the two samples exhibited broad and strong bands at 3426 cm−1 and 1635 cm−1, attributable to H–O–H stretching vibration and twisting vibration of water molecules, respectively, indicating the presence of physically-sorbed interstitial water molecules on the MMIM.41 The peak at 1015 cm−1 corresponds to the bending vibration of the hydroxyl group (metal–OH),42 which was responsible for the formation of inner-sphere surface complexes.43 Two adsorption peaks at 670 cm−1 and 585 cm−1 can be attributed to Fe–O–Fe and Fe–O stretching modes, respectively.18,44 A new adsorption peak at 820 cm−1 appeared on the spectra of As(V)-loaded MMIM, which was attributed to the stretching vibrations of Fe–As–O bond,2,16 while the peaks at 1015 cm−1 significantly weakened, indicating that some hydroxyl groups on the surface of MMIM were replaced by arsenate species. So it could be speculated that the substitution of metal–OH groups by arsenic ions played a key role in the adsorption process.
X-ray photoelectron spectroscopy (XPS) is a versatile analysis technique that was used to investigate the composition and chemical state of MMIM before and after arsenic adsorption. Fig. 4a shows the full-range survey spectra of MMIM and As(V)-loaded MMIM, which indicated that the major elements of MMIM were iron (Fe), manganese (Mn), oxygen (O) and carbon (C). The C 1s spectra before and after arsenic adsorption are presented in Fig. S4,† the peaks at 284.8 eV and 288.6 eV can be assigned to carbon dioxide and carbonate, respectively.45,46 The carbon appearing on the spectra was due to exposure to air and water during sample preparation and reaction. The As 3d spectra on the full-range survey spectra of As-loaded MMIM revealed that the arsenate ions present in aqueous solutions had been adsorbed on the surface of MMIM. The O 1s narrow scans could be deconvoluted into three overlapped peaks corresponding to oxide oxygen (O2−), hydroxyl groups (OH−) and adsorbed water (H2O). From Fig. 4b and c and Table 2, it was found that the O 1s spectra were quite different before and after adsorption. M–O (where M represents metal oxide substrate) increased from 44.504% to 48.535%, this slight increase may be attributed to the new oxygen obtained from As–O after arsenate was adsorbed on the surface of MMIM. However, the percentage of OH− decreased from 36.789% to 32.821%. Arsenic was mainly removed through the hydroxyl groups (OH−) of the M–OH,18 and the high concentration of M–OH was proposed as the main reason for the high arsenic adsorption capacity of MMIM. However, by increasing the arsenic adsorbed on the surface of MMIM, some OH− would be replaced by As(V), so the percentage of OH− decreased after As-loaded MMIM. In was known that the binding energies of the different chemical states of the As 3d core level for As(III) and As(V) are 44.3–44.5 and 45.2–45.6 eV, respectively.47,48 On the spectra of As(V)-loaded MMIM based on As 3d high-resolution deconvoluted spectra (Fig. 4d), only one As(V) 3d peak was detected at 45.58 eV, attributable to As(V)–O bonding, indicating that As(V) has been adsorbed on the surface of MMIM.
![]() | ||
Fig. 4 Full-range XPS spectra of MMIM (a), O 1s spectra with three deconvolutions of MMIM (b), As(V)-loaded MMIM (c) and As 3d core levels of As(V)-loaded MMIM (d). |
Samples | Chemical states | Binding energy (eV) | Percent (%) |
---|---|---|---|
MMIM | O2− | 529.25 | 44.504 |
OH− | 530.77 | 36.789 | |
H2O | 531.57 | 18.707 | |
As(V)-loaded MMIM | O2− | 529.84 | 48.535 |
OH− | 531.41 | 32.821 | |
H2O | 532.24 | 18.644 |
The Fe 2p and Mn 2p XPS spectra are shown in Fig. S5;† the binding energies at 711.29 eV and 724.85 eV were assigned to Fe 2p3/2 and Fe 2p1/2 in Fig. S5a,† and the separation of the 2p3/2 and 2p1/2 spin–orbit levels was approximately 13.6 eV. It was in agreement with the literature that the peaks shifted to the higher binding energy and broaden for Fe3O4.49 In addition, the absence of the satellite peak situated at approximately 719 eV, which was identified as a major characteristic of γ-Fe2O3, also indicated that the iron oxides in MMIM was Fe3O4, because the satellite peak in Fe3O4 would become less resolved or absent.50 These results were in line with the result of WXRD pattern in Fig. 2b. The binding energies of MMIM at 642.2 eV and 653.9 eV were assigned to Mn 2p3/2 and Mn 2p1/2 in Fig. S5b,† indicating the presence of Mn(II), Mn(III) and Mn(IV) states of manganese,51 because the binding energies of Mn(II), Mn(III) and Mn(IV) were very closed to each other.52 This result indicated that although in high purity nitrogen atmosphere, part of Mn(II) was still oxidized into Mn(III) and Mn(IV) during the process of calcinations.
![]() | (1) |
![]() | (2) |
Fig. 5 shows the Langmuir and Freundlich isotherms for arsenate adsorption, and the calculated parameters are summarized in Table 3. It was found that the adsorption data of As(V) were better fitted by the Freundlich model from the correlation coefficients (R2). The maximum adsorption capacities of As(V) were 35.35, 18.85 and 10.86 mg g−1, respectively, indicating that the obtained MMIM was higher uptake of As(V) than that of the template Fe3O4 and the commercial Fe3O4 nanopowders. Compared with commercial Fe3O4 (The BET surface area and pore volume of this commercial Fe3O4 nanopowders were 45.63 m2 g−1 and 0.213 cm3 g−1, respectively), this enhancement in As(V) adsorption capacity was possibly due to the high specific surface area, large pore volume (143.88 m2 g−1 and 0.647 cm3 g−1, respectively) and internal uniform mesopores of MMIM, which can enhance the accessibility of arsenic species to the active sites. However, the BET surface area and pore volume of templated Fe3O4 were 154.36 m2 g−1 and 0.593 cm3 g−1, respectively. But the calculated maximum adsorption capacities of MMIM for As(V) was still much higher than the values of template Fe3O4. This result indicated that this magnetic mesoporous iron manganese bimetal oxides (MMIM) contained iron and manganese oxides presents the synergistic effect than monometallic iron oxides. Therefore, compared with commercial Fe3O4 and template Fe3O4, this obtained MMIM possesses the superiority of bimetal oxides and mesoporous materials, which not only can enhance arsenic adsorption capacity but also presence of the obviously synergistic effect than monometallic iron oxides (Fe3O4).
Samples | Langmuir isotherm model | Freundlich isotherm model | ||||
---|---|---|---|---|---|---|
KL (L mg−1) | qm (mg g−1) | R2 | n | KF | R2 | |
MMIM | 0.643 | 35.35 | 0.9307 | 4.102 | 15.659 | 0.9836 |
Templated Fe3O4 | 1.565 | 18.85 | 0.9137 | 4.029 | 9.986 | 0.9927 |
Commercial Fe3O4 | 6.263 | 10.86 | 0.9128 | 5.489 | 7.320 | 0.9836 |
The results are also significantly higher than that of other reported related adsorbents, such as iron nanomaterials,53 Magnetite–graphene hybrids,1 Fe–BTC polymer,54 γ-Fe2O3 flowers,30 Iron oxide@carbon21 and cellulose@Fe2O3 composites.22 (Table S1†) The constant KF values, which is defined as the arsenic adsorbed on the adsorbent at unit equilibrium concentration. The KF values of MMIM, templated Fe3O4 and commercial Fe3O4 were 15.659, 9.986 and 7.320, respectively, indicating that MMIM exhibited higher adsorption capacity of As(V) than that of templated Fe3O4 and commercial Fe3O4 nanopowders, which agrees well with the conclusions drawn from the Langmuir model. The value of n from MMIM in Freundlich model for As(V) was greater than 1, suggesting this adsorption equilibrium isotherm is nonlinear, which can be attributed to adsorption site heterogeneity, electrostatic attraction and other sorbent–sorbate interactions.55
The adsorption kinetics of As(V) are shown in Fig. 6a. The process is time dependent and the adsorption process can be divided into two steps: (i) the adsorption rate was considerably fast within the first 6 h, which may be due to the large number of available active sites on the surface of MMIM; (ii) the adsorption rate was slow in the subsequent step because the concentration gradients gradually reduce due to the accumulation of As(V) anions on the surface of MMIM, decreasing the adsorption rate of the second stage, and equilibrium capacity was reached within 24 h. This longer equilibrium time may indicate that specific adsorption occurred between the arsenic species and the surface of the adsorbent.56 The pseudo-first order eqn (3) and pseudo-second order eqn (4) were applied to fit the experimental data:
![]() | (3) |
![]() | (4) |
Anion species | Initial concentration (mg L−1) | qe,exp (mg g−1) | Pseudo-first order | Pseudo-second order | ||||
---|---|---|---|---|---|---|---|---|
K1 (h−1) | qe,cal (mg g−1) | R2 | K2 (g mg−1 h−1) | qe,cal (mg g−1) | R2 | |||
As(V) | 5 | 17.255 | 0.1765 | 10.646 | 0.9786 | 0.0495 | 17.699 | 0.9991 |
10 | 26.110 | 0.2283 | 18.015 | 0.9878 | 0.0338 | 26.774 | 0.9989 |
Since the above general kinetic models could not identify the rate-limiting step of As(V) diffusion on MMIM, an intra-particle diffusion model based on the theory proposed by Weber and Morris was used to analyze the rate-limiting step of adsorption,57 as follows:
qt = Kit1/2 + Ci | (5) |
![]() | ||
Fig. 7 Effect of pH on arsenate removal by MMIM at 25 °C. The initial As(V) concentration was 5 mg L−1 and the dosage was 0.2 g L−1. Inset: zeta potential of MMIM and As-loaded MMIM. |
![]() | ||
Fig. 8 Regeneration reuse of the MMIM. The initial As(V) concentration was 5 mg L−1 and the dosage was 0.2 g L−1. |
Based on the above analysis, the adsorption of As(V) on the MMIM in the present study followed a complex mechanism. As(V) removal efficiency by MMIM decreased with increasing solution pH values, indicating that As(V) adsorption by MMIM was not only through ligand exchange under this acidic solution, but also through Coulomb forces. These results suggested that both electrostatic attraction and surface complexation were the major mechanism for the As(V) removal, in addition, the hydroxyl groups on the surface of MMIM also directly exchanged with As(V) to coordinate to structural Fe(III).62
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra09746g |
This journal is © The Royal Society of Chemistry 2015 |