Cation Exchange Superparamagnetic Al-Based Metal Organic

Jan 29, 2018 - ... Marine Living Science, Ocean Science Research Center, Iranian National Institute for ... A cation-exchange metal–organic framewor...
4 downloads 0 Views 3MB Size
Research Article Cite This: ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

pubs.acs.org/journal/ascecg

Cation Exchange Superparamagnetic Al-Based Metal Organic Framework (Fe3O4/MIL-96(Al)) for High Efficient Removal of Pb(II) from Aqueous Solutions Ali Mehdinia,*,† Davoud Jahedi Vaighan,‡ and Ali Jabbari‡ †

Department of Marine Living Science, Ocean Science Research Center, Iranian National Institute for Oceanography and Atmospheric Science, P.O. Box 14155-4781, Tehran, Iran ‡ Department of Chemistry, Faculty of Science, K. N. Toosi University of Technology, P.O. Box 15875-4416, Tehran, Iran S Supporting Information *

ABSTRACT: A cation-exchange metal−organic framework sorbent with high adsorption capacity, active anionic surface, available adsorptive sites, good response in magnetic field and also available enclosed space between the particles for encapsulation of the analyte was in situ prepared by magnetization of a metal organic framework structure of aluminum named as MIL-96(Al). The units of magnetic MIL96(Al) was synthesized by embedding method under hydrothermal condition and well characterized. The mechanism of sorption can be the electrostatic interaction between the anionic structures of the adsorbent with hydrated Pb2+ ions. The high percent of OH groups on the surface of Fe3O4/MIL96(Al) led to high adsorption capacity (>301.5 mg g−1) of the sorbent for Pb2+ ions in aqueous media. The thermodynamic study of Pb2+ adsorption onto the sorbent surface demonstrated that the process was spontaneous, exothermic and physical. The face-centered central composite design was applied for optimizing the adsorption conditions, and the effective parameters were obtained as pH of 6.92, amount of sorbent of 32.4 mg and adsorption time of 25 min. KEYWORDS: Fe3O4/MIL-96(Al), Magnetic composite of MOF, Removal of lead ions, Magnetic metal organic trap, Cation-exchange MFCs, Central composite design



INTRODUCTION Metal−organic frameworks (MOFs) or porous coordination polymers, consisting metal ions/clusters and organic struts, have been recognized as an excellent platform for host−guest chemistry applications1 such as gas storage,2 ion-exchange,1 separation,3 catalysis,4 drug delivery,5 sensors6 and conduction.7 The main advantage of MOFs in comparison to the alternative materials is the tunability of pore size, shape, topology and functionality.8 Selective sensing and exchange of ions is one of the main application of porous materials.9 Recently, ionexchange MOFs are new category of MOFs with large surface area, high porosity and charged framework which led to their high efficiency for ion-exchange chromatography and ionexchange solid-phase extraction. The strong columbic interaction between the charged stationary phase and charged analytes leads to high efficiency of separation process.1 In addition, ion-exchange MOFs, due to their large surface area and porosity, have well capacity for solid-phase extraction of the charged analytes. The composition of MOFs and the materials with a variety of functional groups is performed to combine the merits and reduce the shortcomings of the both components.10 MOFs of © 2018 American Chemical Society

highly porous nanocomposites were made with active nanoparticles such as metal nanoparticles/nanorods,11,12 oxides,13,14 quantum dots,15,16 polymers,17,18 polyoxometalates,19 graphene,20 carbon nanotubes,21,22 biomolecules23 and so on. In this regard, magnetic framework composites (MFCs), fabricated by combination of magnetic nano- or microparticles with MOFs crystals,24 are good candidates for the adsorption process. Several methods introduced for fabrication of MFCs such as embedding,25 encapsulation,26 mixing27 and layer-bylayer28 strategies. MFCs combine the favorable virtues of both magnetic Fe3O4 nanoparticles and metal organic framework. It makes them excellent candidates for adsorption process,27 and also the phase separation can be rapidly and easily accomplished by applying an external magnetic field.29 The 3D metal−organic framework of MIL-96 (Al) (MIL stands for Material of Institute Lavoisier), with unit cell of Al12O(OH)18(H2O)3(Al2(OH)4)[btc]6·24H2O, involves the aluminum octahedral units interacting with the btc ligands. It Received: September 16, 2017 Revised: January 22, 2018 Published: January 29, 2018 3176

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering was hydrothermally synthesized for the first time by Loiseau and co-workers and applied for separation of CH4, CO2 and H2.30 In the other works, MIL-96 (Al) was used for separation of C5-hydrocarbons,31 vapor-phase adsorption of alkylaromatics,32 adsorption of nitrogenous volatile organic compounds (VOCs)33 and defluorination of drinking water.34 The MIL-96 (Al) is a practical material due to using of inexpensive metal of aluminum,35 high hydrothermal stability,36,37 flexible structure,32 excellent stability in neutral-to-acidic aqueous solutions38 and high porosity.30 Furthermore, existence of large number of hydroxyl groups in its structure makes it useful for extraction purpose. Also, the weak Lewis acid sites in the AlMIL-96 structure could form a coordinated bond with water molecules, hydroxyl groups or other hard base groups in the aqueous solution.33 It can also create negative charges in the structure of MIL-96 (Al). Therefore, MIL-96 (Al), due to its interaction by water molecules, has high potential for release of protons and well capacity for selective cation-exchange application. In the neutral pH, O− groups in the framework of MOF lead to form anionic structure for it. Hence, MIL96(Al) possesses several effective factors to enhance efficiency of the extraction process. First, combining the specific pore size and strong columbic interaction between the anionic framework of this MOF and cationic compounds can cause selective extraction of the cationic analytes.1 Second, according to the Pearson’s hard/soft acid/base concept,39 hard base oxygen atoms in the framework of MOF can form stable complexes with the analytes with hard acids groups. Third, the replacement of water molecules in the hydrated structure of cationic metals by oxygen atoms in the framework of MOFs40 can effect on the selectivity of MOF. Fourth, spontaneous aggregation of MIL-96(Al) nanoparticles, during extraction process, can create a trap with the available enclosed space between the particles for capsulation of the bulky compounds. The aggregation can occur due to the interaction of hard Lewis acid sites of Al in MIL-96 particles33 with surficial oxygen atoms in the adjacent MIL-96(Al) particles. It also leads to further increasing in sorption capacity of the sorbent and its ability for separating of the cloudy precipitated metals in aqueous solution. Also, it can cause rapid and easily isolation of the sorbent from aqueous solution after sorption procedure by an external magnetic field. In the present work, the MIL-96 (Al) was considered for design of a magnetic cation-exchange metal−organic trap with the advantages of ion-exchange MOFs and MFCs. Also, synthesis of a magnetic nanocomposite of MIL-96(Al) was illustrated and its potential ability for in situ preparation of cation-exchange metal−organic trap was investigated. It was used for adsorption of lead(II) ions in aqueous solution by trapping and cation-exchange mechanism. Parameters affecting the efficiency of sorption were evaluated in detail by central composite design under response surface methodology (RSM).



ization of Fe3O4/MIL-96(Al). Sodium hydroxide pellets (NaOH) (Merck, 99%) and nitric acid supra pure (HNO3) (Sigma-Aldrich, 65%) were required for the adsorption and adjusting pH of the solutions. Stock standard solution (1000 mg L−1) of Pb2+ was prepared by dissolving of 0.159 g of analytical grade Pb(NO3)2 salt (Merck) in 2 mL of 6N nitric acid and dilute to 100 mL by the deionized water. Working standard solutions were prepared daily by diluting the stock standard solution to the appropriate concentrations with the deionized water. Apparatus. The crystalline structure and composition of the Fe3O4/MIL-96(Al) were identified by Powder X-ray diffraction (PXRD, Philips X’pert diffractometer, model pw1730), field emission scanning electron microscopy (FE-SEM, MIRA3TESCAN-XMU) with EDX analysis, scanning electron microscopy (SEM, KYKYEM3200, −26 kV), transmission electron microscopy (TEM, Philips), Fourier-transform infrared spectroscopy (FT-IR, Bruker, VERTEX 70), vibrating sample magnetometry (VSM, Daghigh Kavir Co., Kashan, Iran) and dynamic Light Scattering (DLS, Malvern, Zetasizer Nano ZS). The surface of the of the synthesis Fe3O4/MIL-96(Al) was analyzed by X-ray photoelectron spectroscopy (XPS, Thermo Scientific ESCALAB 250Xi). The nitrogen adsorption/desorption isotherms were obtained at −196 °C with an adsorption unit (BET, Belsorp mini II, Japan Co.) after evacuation at 200 °C for 3 h. The surface area was calculated from nitrogen adsorption isotherms using the Langmuir and BET equations. Digital pH meter, Metrohm 827 (Titrino, Metrohm, Switzerland) was used for pH measurements. The GBC 932 plus (Australia) flame atomic absorption spectrophotometer (FAAS) equipped with a lead hallow cathode lamp (GBC, Australia) at the wavelength of 217 nm was applied for determination of lead ions in the sample solutions and thermal energy created via air/acetylene flame, applied for generation of Pb gas phase atoms in all measurements. Synthesis of Fe3O4 Nanoparticles. Fe3O4 nanoparticles were synthesized via the coprecipitation method;41 8.1 g of iron(III) chloride (FeCl3·6H2O) and 3 g of iron(II) chloride (FeCl2·4H2O) were dissolved in 150 mL of deoxygenated deionized water during stirring and under nitrogen gas atmosphere. 25 mL of NH4OH (25%, v/v) was added to this solution at 80 °C under strong stirring conditions. Stirring of the reacted mixture was continued for 30 min, and then cooled to room temperature. The black Fe3O4 MNPs were isolated by a magnetic field and washed three times with the deionized water. Synthesis of MIL-96(Al). Particles of Al12O(OH)18(H2O)3(Al2(OH)4)[btc]6·24H2O, MIL-96(Al) were prepared via a hydrothermal method that has been reported previously.38,42 First, 750 mg of Al(NO3)3·9H2O was dissolved in 10 mL of water− DMF mixture (4:1), then 420 mg of 1,3,5-benzenetricarboxylic acid (btc) was added to the reaction mixture and dissolved with 2 h of sonication. Afterward, the resulting mixture was refluxed at 140 °C for 24 h. Finally, a white color product was separated by centrifugation and washed three times with deionized water and methanol. Then, the product was dried at 100 °C in an oven for 12 h. Synthesis of Fe3O4/MIL-96(Al). Hydrothermal method was used for embedding24 of Fe3O4 nanoparticles in MIL-96(Al) crystals. Initially, 2.25 g of Al(NO3)3·9H2O was dissolved in a 30 mL mixture of water:DMF (4:1). Then, 0.9 g of Fe3O4 nanoparticles, synthesized in the previous step, was dispersed in the reaction mixture under sonication (15 min). Next, 1.26 g of H3BTC was added to the reaction flask and sonicated in an ultrasonic bath for 2 h under mechanical stirring. Resulting suspension was refluxed at 140 °C for 36 h. Produced nonmagnetic composite was separated by an external magnet and washed three times with deionized water. At the end, the product was washed three times by methanol and dried at 100 °C in an oven for 12 h. Phosphate Functionalization of Fe3O4/MIL-96(Al). 100 mg of NaH2PO4·H2O was dissolved in 20 mL of deionized water, and then 100 mg of the prepared Fe3O4/MIL-96(Al) was added to this solution, the pH was set at 6.5 and the mixture stirred for 12 h at 25 °C. Subsequently, the obtained product was isolated from the reaction

EXPERIMENTAL SECTION

Materials and Solutions. The reagents were obtained from a commercial seller and used without further purification. Benzene-1,3,5tricarboxylic acid (H3BTC) (Sigma-Aldrich, 99%), aluminum nitrate nonahydrate (Al(NO3)3·9H2O) (Merck, 98.5%), dimethylformamide (DMF) (Sigma-Aldrich, 99.8%), methanol (Sigma-Aldrich, 99.8%), iron(III) chloride hexahydrate (FeCl3·6H2O) (Merck, 99%), iron(II) chloride tetrahydrate (FeCl2·4H2O) (Merck, 99%) and ammonium hydroxide solution (NH4OH) (Merck 25%) were used for synthesis of Fe3O4/MIL-96(Al) nanocomposite. Sodium phosphate monohydrate (NaH2PO4·H2O) was obtained from Merck and used for functional3177

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering

Figure 1. Schematic illustration of two possible mechanisms for embedding Fe3O4 nanoparticles into the MIL-96(Al) crystals (a). The SEM images of Fe3O4 nanoparticles (b) and MIL-96(Al) nanoparticles (c), FE-SEM image Fe3O4/MIL-96(Al) composite (d) TEM images of Fe3O4/MIL-96(Al) composite (f, g and q).



mixture by external magnet, washed three times by deionized water and methanol. Finally, the product was dried at 100 °C in an oven. Batch Cation Exchange Experiments. The procedure of magnetic cation exchange adsorption of lead ions was performed in batch experiments. According to a preliminary experimental design, 32.4 mg of Fe3O4/MIL-96(Al) was activated (Figure S1) via dispersing in 3 mL of 1 M NaOH solution by 20 s of sonication in an ultrasonic bath, and after 10 min the activated sorbent was collected by a permanent magnet (2 cm × 4 cm × 6 cm, 1.3 T) and washed by deionized water. Then, it was added in a 100 mL aqueous sample containing the lead ions, and pH was adjusted to 6.92 by the diluted solution of NaOH or HNO3. The solution was stirred softly for 25 min at room temperature (25 ± 1 °C) using a magnetic stirrer at a steady rate of 750 rpm to suspend MFCs thoroughly and facilitate adsorption lead ions onto the sorbent. At the end of the extraction time, the solution was transferred to the conical tubes and after collecting the sorbent by a magnet on one side of the tube, the supernatant was decanted. The amounts of Pb2+ ions were specified using FAAS. Experimental Design Methodology. RSM is an alternate statistical approach to obtain the effect of the experimental variables that can mainly effect on the extraction process.43 The exclusive factors must be considered along with nonlinear effects and interaction terms.44 Central composite design (CCD), appropriate for fitting a quadratic surface, was used for optimization of the effective parameters with a minimum number of experiments and also for analyzing of the interaction between the parameters.45 In this case, face-centered central composite design (FC-CCD) factors were assessed at three levels and the axial points were set on the centers of faces of the cubic experimental area. As a result, center point replications in FC-CCDs are generally less than the amount required for the comparable rotatable CCD. So, it can reduce the excessive changing of factor levels, makes comfortable administration of the experiment, and also can decrease the required time and expenses.46 For this reason, the FC-CCD was used for optimization of the adsorption process of Pb2+ ions and elution step. Then, Design-Expert 7.1.3 software was applied for analysis of the experimental design data and estimation of the predicted responses.

RESULTS AND DISCUSSION Characterization of the Fe3O4/MIL-96(Al). In a typical synthesis, embedding method was used for synthesis of Fe3O4/ MIL-96(Al) in the hydrothermal condition. The magnetic Fe3O4 nanoparticles were added into the mixture solution of ligands and inorganic precursors of MIL-96(Al). Two mechanisms can be offered for embedding of Fe3O4 nanoparticles within the MIL-96(Al) crystals (Figure 1a). In a possible mechanism, according to this fact that Al(III) is a hard Lewis acid39 with empty d orbitals, and aluminum octahedral units is a label complex,35 they can form a stable coordination bound with OH hard base groups on the surface of Fe3O4 nanoparticles. Then, nucleation can carry out around of the Fe3O4 nanoparticles. In another possible mechanism, H3BTC ligand can be coordinated with Fe(III) sites on the surface of Fe3O4 nanoparticles. Subsequently, MIL-96(Al) crystals, after the heat-driven process, can grow in the surrounding of Fe3O4 nanoparticles. To determine the surface chemical composition of the synthesized Fe3O4/MIL-96(Al) and evaluate the probable connection mechanism between Fe3O4 and MIL96(Al) X-ray photoelectron spectrometry results are shown in Figure 2. The wide scan spectrum of Fe3O4/MIL-96(Al), as shown in Figure 2a, indicated the presence of aluminum, carbon, oxygen, and iron in the composite. The photoelectron lines at binding energies of approximately 75, 285, 530 and 710 eV are attributed to Al 2p, C 1s, O 1s and Fe 2p, respectively. The high-resolution XPS spectrum for O 1s is shown in Figure 2b. In the O 1s spectrum of Fe3O4/MIL-96(Al), there were different oxygen containing groups including metal oxides (530.4 eV), Fe−O−C (532.1 eV), Al−OH and aluminum carbonate (532.3 eV) and the peak at 533.7 eV should correspond to C−O groups.47,48 It is worth mentioning that the Fe−O−C peak at 532.1 eV was ascribed to some covalent 3178

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering

Figure 3. (a and b) Simulated XRD patterns of MIL-96(Al) and Fe3O4, extracted from a cif-file presented by Loiseau and co-workers30 and cif-file presented by O’Neill and co-workers,49 respectively. (c) Experimental XRD pattern of Fe3O4. (d) XRD pattern of MIL-96(Al). (e) XRD pattern of Fe3O4/MIL-96(Al).

Figure 2. (a) XPS survey spectra of Fe3O4/MIL-96(Al). (b) Fitted results of XPS O 1s spectra.

XRD pattern of Fe3O4/MIL-96(Al) after activation by NaOH solution was prepared (Figure S3). Comparison of the XRD pattern of Fe3O4/MIL-96(Al) before and after activation indicates the appropriate stability of Fe3O4/MIL-96(Al) during the activation process. The N2 adsorption−desorption isotherm plot was exhibited for Fe3O4/MIL-96(Al) in Figure S4. The Langmuir and BET surface areas, calculated from the adsorption isotherm, were around 98.3 and 73.8 m2 g−1, respectively. BET surface areas for MIL-96(Al) was 216 m2 g−1 in the literature.36 It can prove that the Fe3O4 nanoparticles were embedded in the MIL-96(Al) crystals and reduced the effective surface area of MIL-96(Al). The FT-IR spectra of Fe3O4, MIL-96(Al) and Fe3O4/MIL96(Al) are shown in Figure S5a,b,c, respectively. The two bands observed at 632 and 585 cm−1 in Figure S5a are as a result of splitting of the FeO absorption band of bulk Fe3O4 at 570 cm−1 that shifted to the higher wavenumber.50 The characteristic absorption bands of Fe3O4 at 632 and 585 cm−1, observed in Figure S5c, indicated successful embedding of magnetic nanoparticles in the MOFs crystals. The FT-IR spectra in Figure S5b indicated conformity of the characteristic absorption bands of the synthesized MIL-96(Al). It confirmed with the FT-IR spectra obtained by Loiseau et al.30 The band at 625 cm−1 is assigned to (OH)AlO angle bending of aluminum cluster51 (Figure S5b). The vibrational bands in the range of 1400−1600 cm−1 at the FT-IR spectra in Figure S5b,c, are related to aromatic rings in the ligand groups of MIL-96(Al). Also, there are asymmetric stretching vibrations of bridging carboxylate groups in the range of 1500−1700 cm−152 and the symmetric vibrations between 1459 and 1399 cm−1, which is assigned to the CO bond.30 Due to the fully coordination of aluminum units with the btc ligands, strong absorption bond of carboxylates CO groups in the area of 1710−1740 cm−1 disappeared. Figure 4 shows the magnetization curves (VSM curve) of Fe3O4 and Fe3O4/MIL-96(Al). As can be observed from the figure, both samples exhibited a superparamagnetic property. The saturation magnetization of Fe3O4 and Fe3O4/MIL-96(Al)

bonds between Fe3O4 and BTC ligand (Figure 2b). H3BTC ligand can be coordinated with Fe(III) sites on the surface of Fe3O4 nanoparticles. Subsequently, MIL-96(Al) crystals, after the heat-driven process, can grow in the surrounding of Fe3O4 nanoparticles. That is corresponded with proposed mechanism for synthesis of Fe 3O 4/MIL-96(Al), and confirms the connection between Fe3O4 and MIL-96(Al) through oxygen atoms in carboxylates groups of BTC ligands and Fe3+ in Fe3O4. Figure 1b shows the SEM image of pure Fe3O4 with granular structure. The spindle shaped particles, obtained from original MIL-96(Al) is shown in Figure 1c. It is in accordance with the proposed morphology for MIL-96(Al) in the literature.38 Typically, the polycrystalline Fe3O4/MIL-96(Al) particles tend to form crystal morphology similar to the original MIL96(Al). FE-SEM and TEM images (Figure 1d,f,g,q) of Fe3O4/ MIL-96(Al) almost indicated these similarities in crystal morphology. Also, TEM images (Figure 1f,g,q) show that the Fe3O4 nanoparticles were dispersed in the MIL-96(Al) crystals. In addition, the composition of the sorbent was determined using EDS analysis as shown in Figure S2. The results confirmed the presence of elements of the Fe3O4 nanoparticles distributed in the structure of Fe3O4/MIL-96(Al) composite. XRD analysis of the sorbent is indicated in Figure 3. The peaks in the XRD analysis of Fe3O4/MIL-96(Al) can be related to crystalline Fe3O4 and MIL-96(Al) (Figure 3e). The diffraction peaks observed at ∼30.0°, 35.4°, 43.0°, 53.4°, 56.9° and 62.5° (Figure 3e) can be attributed to the typical cubic spinel structure of Fe3O4 which embedded in MIL96(Al). Also, they have a great conformity with the Fe3O4 XRD pattern simulated by O’Neill et al.49 (Figure 3b) and experimental XRD pattern of Fe3O4 (Figure 3c). The patterns (Figure 3d,e) illustrate that the diffraction peaks at ∼5.651°, 7.72°, 9.14°, 13.65°, 14.63°, 15.49°, 16.7°, in the magnetic Fe3O4/MIL-96(Al) and MIL-96(Al) can be matched with the diffraction peaks of simulated MIL-96(Al), that presented by Loiseau and co-workers30 (Figure 3a). It indicates the successful synthesis of Fe3O4/MIL-96(Al) composite. Also, 3179

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering

Lewis acid behavior of the sorbent was studied by titration. 50 mL of distilled water suspension containing 50 mg of Fe3O4/ MIL-96(Al) was titrated by 0.1 M NaOH solution. The titration plot (Figure 5a) indicated that the sorbent has high potential for release of proton in aqueous solution due to the presence of Lewis acid sites of Al on the sorbent. The coordination of water molecules (as hard base) with Al atoms (as hard Lewis acid) leads to release of proton of the coordinated water molecules, and creates anionic frameworks with high coordinated hydroxyl groups on the surface of Fe3O4/MIL-96(Al) particles in aqueous solution. Also, ζpotential measurement, presented in Figure 5b, shows the surface charge of the Fe3O4/MIL-96(Al) particles at different pH values. The isoelectric point (IEP) of Fe3O4/MIL-96(Al) particles was found to be 5.13, and it confirms anionic nature of Fe3O4/MIL-96(Al) in neutral aqueous solution. Evaluation of Adsorption of Different Metal Ions on the Surface of Fe3O4/MIL-96(Al). Different hydrated metals ions; hard, intermediate and soft metal ions, with different labilities, were investigated to study the adsorption mechanism. As presented in Table 1, there are several effective factors in the

Figure 4. VSM magnetization curve of Fe3O4 (a), Fe3O4/MIL-96(Al) (b).

were obtained as 54.5 and 24.6 emu g−1, respectively. Due to the embedding of Fe3O4 nanoparticles in the crystalin structure of MIL-96(Al), the obtained superparamagnetic Fe3O4/MIL96(Al) particles showed enough magnetism for facile separation from sample matrix by applying a magnetic field. As can be seen from Figure 4, the saturation magnetization was gradually decreased with increasing MOFs contents on the superparamagnetic nanoparticles.53,54 The obtained values of saturation magnetization can be used to estimate the weight percentage (w/w %) of magnetic Fe3O4 nanoparticles (Fe3O4 %) and MOF contents (MIL96(Al) %) in the particles of Fe3O4/MIL-96(Al). Equations 1 and 2 were proposed for estimation of the composition of each nanoparticle: Fe3O4 % =

SMFe3O4 /MIL‐96(Al) SMFe3O4

MIL(96)% = 100 − Fe3O4 %

Table 1. Hard/Soft and Labile/Inert Effects on the Extraction Efficiency of Different Metals Ions on the Surface of Fe3O4/MIL-96(Al) and MIL-96(Al)a Extraction % Metals

Hard/Soft39

kH2O (s−1)55,56

Fe3O4/MIL-96(Al)

MIL96(Al)

Cd2+

Soft

108 < kH2O < 109

0

0

2+

Soft

109 < kH2O < 1010

66 ± 3

69 ± 2

Cu2+

Borderline

109 < kH2O < 1010

98 ± 2

97 ± 3

(1)

Pb2+ Ni2+

Borderline Borderline

Labile 104 < kH2O < 105

92 ± 1 0

90 ± 3 1±1

(2)

Zn2+

Hg

× 100

Whereas SMFe3O4 and SMFe3O4/MIL‑96(Al) are saturation magnetizations of pure Fe3O4 nanoparticles and Fe3O4/MIL-96(Al) particles, respectively. For synthesized magnetic composite, calculated Fe3O4 % and MIL96(Al) % were 45.14% and 54.86%, respectively. These data are an estimate of MOF growth around of the Fe3O4 nanoparticles for each synthesis. Evaluation of Chemical Characteristics of Fe3O4/MIL96 (Al). Study and comparison of the adsorption amount of different analytes on the surface of Fe3O4/MIL-96(Al) demonstrated the appropriative properties of the sorbent.

a

Borderline

107 < kH2O < 108

5±1

3±1

3+

Cr Na+

Hard Hard

Inert 108 < kH2O < 109

40 ± 3 −

44 ± 2 −

Al3+

Hard

kH2O ∼ 10°

Aggregation of sorbent particles



(kH2O) is water exchange rate constants.55,56

ion exchange adsorption process: content of negative charges on the surface of adsorbent, quantity of positive charges on the metals ions, amount of basic groups with different hardness on the surface of sorbent, hardness extent of metal ions and kinetic

Figure 5. Titration curves of 50 mg Fe3O4/MIL-96(Al) by NaOH 0.1 M (a), ζ-potential as a function of the pH for Fe3O4/MIL-96(Al) (b). 3180

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering of water exchange in metal ions.40,55−57 For cation exchange sorption of metal ions by Fe3O4/MIL-96(Al), amounts of negative and positive charges on the surface pores of the sorbent and metal ions, respectively, produce the required force for immigration of metal ions from the bulk of solution onto the cavity surfaces of the adsorbent. This electrostatic force increases immigration speed of the metal ions, but this electrostatic force is not enough to provide required selectivity for the extraction of each metal ion. In this regard, two other factors can affect to obtain the stable and selective extraction. In the second sorption stage, the metal ions require to be retained by the sorbent after immigration of cationic metals to the surface of sorbent. It is due to the stable interaction of the hydroxyl groups on the surface of sorbent and cationic metal ions. But, these stable interactions may be time-consuming for some ions, because of the need to different activation energies for substitution of the coordinated water molecules with the different hydrated metal ions. Thus, each ion that can quickly enter in the substitution reaction was retained by the sorbent and other ions were exchanged by those ions. The primary experiments (Table 1) showed that the Fe3O4/MIL-96(Al) represents good response for extraction of Cu2+ and Pb2+ ions. Because of the higher environmentally importance of Pb2+ ion, it was used as analyte for subsequent experiments. Extraction recoveries of different cations were also investigated for pure MIL-96(Al). The similar results were obtained for MIL-96(Al). It indicated that Fe3O4 nanoparticles embedded into the MIL96(Al) can compensate the reduction in the BET surface area of Fe3O4/MIL-96(Al) than MIL-96(Al). According to the results, the adsorption mechanism of Pb2+ onto the cavity of Fe3O4/MIL-96(Al) can be proposed according to Figure S6. In the natural and alkaline pH values, the sorbent has negative surface charges. The negative charges of the activated sorbent are neutralized by proton or Na+ cations. After entering the sorbent to the sample solution, the hydrated Pb2+ cations rapidly immigrated58 from the bulk of solution to the surface and cavities of Fe3O4/MIL-96(Al). Charge equilibrium on the surface of the activated Fe3O4/MIL96(Al) was tuned by ejecting Na+ and other metal cations from the surface of the sorbent. Then, due to the suitable labiality of the hydrated Pb2+ ions at room temperature, substitution reaction and sorption step can be rapidly and completely performed. In addition, the aggregation of Fe3O4/MIL-96(Al) was observed slowly during the Pb2+ sorption. Figure S6 also indicated the schematic mechanism of aggregation of the particles. Actually in natural pH, Fe3O4/MIL-96(Al) particles have dual properties. The hydroxyl groups on a particle can interact by Lewis acid Al sites on the adjacent particle. Al3+ is located in the main groups of metal ions with high positive charge density and empty d orbitals, hence it can interact successfully by hard bases.56 But due to the high positive charge density and low ionic radius, Al3+ has stable56 interaction by the hydroxyl or water molecule in aqueous solution. Therefore, Al3+ has low lability compared to Pb2+ ions,56 and aggregation of the particles occurs slower than adsorption of Pb2+ ions in the studied temperature. It helps the extraction process, because after completing the extraction of Pb2+, the aggregation of sorbent particles is completed. Thus, the aggregation has a low effect on the extraction efficiency due to the decreasing in dispersity of the sorbent in the batch sorption process. In addition, the data obtained from the capacity measurement (Figure 6) indicated that the capacity of the sorbent highly enhanced by trapping the precipitated analyte during

Figure 6. Plot of Qe (mg g−1) versus initial concentration of Pb2+ (mg L−1) at temperature of 25 °C and pH of 6.92.

aggregation of Fe3O4/MIL-96(Al) particles. Other benefits of magnetic sorbent particles aggregation included easy magnetic isolation of these particles after batch sorption stage. By aggregation of the nanoparticles, the lumped particles will have lower surface area than the dispersed sorbent particles, and a reduction can be occurred in friction force between solvent molecule and particles surfaces. Thus, aggregation after sorption stage can reduce the required magnetic force for isolation of the sorbent. The EDS analysis of Fe3O4/MIL96(Al) after adsorption of Pb2+ was also performed to prove Pb2+ adsorption. The obtained results (Figure S7) indicated the presence of Lead ions in the surface of Fe3O4/MIL-96(Al) after adsorption process. Dual Behavior of Fe3O4/MIL-96(Al). Due to the Lewis acid sites, MIL-96(Al) is recently used for adsorption of nitrogenous volatile organic compounds.33 For further study on activity of Lewis acid sites of Al on the surface of Fe3O4/MIL96(Al), phosphate ions were used as model compounds. Phosphate ions, having hard base groups could creative effective interaction with Lewis acid sites of Al on the surface of Fe3O4/ MIL-96(Al). The FT-IR spectra (Figure S8) of the phosphate functionalized Fe3O4/MIL-96(Al) indicated the capability of Lewis acid sites of Fe3O4/MIL-96(Al) for establishing stable interaction with the phosphate ions. The band obtained at 1000−1200 cm−1 for coordinated phosphate confirmed the stable interaction of phosphate with the surface sites of the sorbent. The active Lewis acid sites of Fe3O4/MIL-96(Al) can be useful for using of this sorbent for extraction of organic compounds, and facile modification of Fe3O4/MIL-96(Al) for extraction of other metal cations. Optimization of Extraction Parameters. Face-centered central composite design (FC-CCD), by three times replicate center point, was applied for optimization of the sorption step and estimation of the interaction of effective parameters including pH of solution (A), sorption time (B) and amount of magnetic sorbent (C). The extraction percentage was considered for the responses of FC-CCD. The levels of factors and design matrix with responses are shown in Tables S1 and S2. Experiments were distributed randomly to minimize the effects of uncontrolled variables. Distance center point from axial point was introduced by α value,45 and adjusted (α = 1) for FC-CCD. After analysis of variance (ANOVA) (Table S3), response was modeled by the following quadratic polynomial equation: 3181

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering

Where C0 and Ce (mg L−1) are the primary and equilibrium concentrations of Pb2+ in solution, M (g) is the mass of the adsorbent and V (L) is the volume of solution. The results, presented in Figure 6, indicated that the amount of Qe increased by increasing the initial concentration of Pb2+ up to 100 mg L−1. The equilibrium sorption capacity (Qe = 301.5 mg g−1) indicated that Fe3O4/MIL-96(Al) has a very good capacity for sorption of Pb2+ ions. Also, the maximum amount of adsorption capacity was not achieved in the study range. This high equilibrium sorption capacity can be attributed to high porosity of the magnetic sorbent. Two strategies can be concluded by comparison of the plots of Qe and Qe/Ce vs Ce in Figure S10. It is indicated that when Ce < 0.43 μg mL−1, the ratio of Qe/Ce was rapidly increased that shows the ion exchange sorption mechanism. But, when Ce > 0.43 μg mL−1, the ratio of Qe/Ce was gradually constant, due to the precipitation of Pb2+. Thus, the isotherm models were studied at low concentrations of Pb2+. Experimental data were analyzed using Freundlich, Temkin and Dubinin−Radushkevich isotherm models. Linearized equations are expressed as below: 1 Freundlich model: ln Q e = ln Ce + ln K f (5) n

Log10(Extraction% + 5.00) = 1.96 + 0.46A + 0.028B + 0.029C − 0.019AB − 0.018AC + 0.012BC − 0.42A2 − 0.020B2

(3)

Where A, B and C are coded values of independent variables. The coefficients with one factor demonstrate the effect of the specific factor, whereas the coefficients with two factors and those with quadratic phrases display the interaction between the two factors and second-order effect, respectively. The positive marks in front of the terms show synergistic effect, whereas negative marks display opposite effect.45 It was found the quadratic model is the best, due to the low standard deviation and high R2 statistics. The values of standard deviation and R2 for the extraction percent of Pb2+ were 0.015 and 0.99, respectively. Coefficient of eq 3 clearly indicated that the pH of solution has the highest significantly positive effect in the adsorption step. The decrease in extraction percent at low pH values (Figure S9a,b and Figure S9c,d) can be attributed to the protonation of active sites of the magnetic sorbent and reducing the negative charges in the structure of sorbent. On the other hand, the low extraction percentages at high pH values can be due to the increasing in aggregation rate of Fe3O4/MIL-96(Al) particles versus sorption rate of Pb2+ at these pH values. Also, interaction plots (Figure S9b,d) indicated the highest antagonism interaction of pH by other variables in very low and high pH values. It can be attributed to decreasing in accessibility of the active sites of Fe3O4/MIL-96(Al) at very low and high pH values. Figure S9g,h represents the positive effects of increasing in active sites and their availability on the extraction efficiency by increasing the amount of sorbent and sorption time. Also, Figure S9h presents the synergic effect of the sorption time and amount on the extraction efficiency, due to enhancement in availability of active sites of sorbent for interaction with Pb2+ ions. Low slope of the curve, in plot of extraction % vs sorption times, indicates rapid immigration of Pb2+ ions toward the sorbent through the electrostatic force between Pb2+ ions and the Fe3O4/MIL-96(Al) surface. In addition, the low difference of extraction percentages in low and high levels of sorbent amount at the constant extraction time and pH, could be the reason for the high porosity and capacity of anionic structure of Fe3O4/MIL-96(Al) for the extraction of Pb2+ ions (Figure S9h). Conforming to the acquired results from the optimization study, the following experimental conditions were chosen: pH, of 6.92; amount of magnetic sorbent of 32.4 mg and extraction time of 25 min. Sorption Capacity Study. One of the important characteristics of the sorbents in SPE is the sorption capacity. The sorption capacity parameter indicated potential ability of the adsorbent for sorption of the target analyte(s). In this regard, 32.4 mg of active adsorbent was added to the solutions (100 mL) with various concentrations of Pb2+ (0−100 mg L−1). After the pH was adjusted at 6.92 and the solution stirred for 25 min, the sorbent was separated by a magnet and the concentration of Pb2+ in the supernatant was determined by FAAS. The equilibrium concentration of Pb2+ ions, Qe (mg g−1) adsorbed per gram of sorbent, was calculated according to the following equation: Qe =

(C0 − Ce)V M

Temkim model: Q e = K t ln Ce + K t ln f

(6) 2

Dubinin−Radushkevich model: ln Q e = ln Q s − (Kdε ) (7)

ε = RT ln(1 + 1/Ce)

(8)

−1

Where Qs (mg g ) is the theoretical saturation capacity. Kf, Kt and Kd are the adsorption constants of Freundlich, Temkin and Dubinin−Radushkevich models, respectively, and n is the Freundlich linearity index. f (L g−1) is the Temkin isotherm equilibrium binding constant. R (8.314 Jmol−1 K−1) is the universal gas constant and T (K) is the absolute temperature. The fitting results based on these three models are listed in Table 2. It appears that adsorbate−adsorbent system can be Table 2. Parameters of Isotherm Models Freundlich isotherm Kf mg/g(L/mg)1/n 19777 Kt (J/mol) 136.75

n 0.18 Temkin isotherm f (L g−1)

4.247 Dubinin−Radushkevich isotherm

R2 0.893 R2 0.994

Qs (mg/g)

Kd (mol2 J−2)

R2

5506

4.047 × 10−7

0.915

better explained by the Temkin model (Figure S11), in accordance with the correlation coefficient (R2 = 0.994) result. The Temkin model is a proper model for adsorption, based on electrostatic interaction between positive and negative charges,59 which is compatible by ion exchange adsorption mechanism in sorption process. Thermodynamics of Adsorption. To evaluate the mechanism of Pb2+ adsorption on the Fe3O4/MIL-96(Al), the extractions were performed at different temperatures (11, 30 and 50 °C) from the solutions containing constant amount of Pb2+ (300 μg). Thermodynamic parameters, the standard

(4) 3182

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering

Reusability of Fe3O4/MIL-96(Al). Reusability of Fe3O4/ MIL-96(Al) is related to the stability of MOF district in the structure of Fe3O4/MIL-96(Al). Many MOFs are unstable in aqueous solution, and it limits their application in aqueous solution, but MIL-96(AL) is one of the stable MOFs in aqueous solution.38 Reusability of Fe3O4/MIL-96(Al) was investigated. Figure S12 shows good extraction efficiency of Pb2+ after two cycles of extraction and regeneration. It shows relatively acceptable stability of Fe3O4/MIL-96(Al) in alkaline and acidic condition in extraction and regeneration process. To represent the application benefit of Fe3O4/MIL-96(Al) in adsorption of Pb2+, a number of adsorbents, recently used in other research, are compared with the new adsorbent in Table 3. The maximum adsorption capacity values in this work were

enthalpy (ΔH°), standard entropy (ΔS°) and Gibbs free energy (ΔG°), for adsorption of Pb2+ on Fe3O4/MIL-96(Al) were calculated by using Van’t Hoff equation (eqs 9 and 10):33 ln

Qe Ce

=

ΔS° ΔH ° − R RT

(9) (10)

ΔG° = ΔH ° − T ΔS°

Where T is the absolute temperature (K) and R is the gas constant (8.314 J mol−1 K−1). ΔH° and ΔS° were obtained as −19.99 kJ mol−1 and −35.91 J mol−1 K−1, from the slop and intercept of the linear curve of ln

Qe Ce

as a function of

1 , T

respectively. ΔG° values were calculated (−9.79, −9.11 and −8.39 kJ mol−1 K−1) at each temperature (11, 30 and 50 °C). The data obtained at various temperatures showed that the extraction percentages gradually decreased with increasing the extraction temperature. These results are compatible with the obtained negative enthalpy (ΔH° < 0), and demonstrated that the adsorption of Pb2+ ions on the Fe3O4/MIL-96(Al) is an exothermic reaction. It is due to this fact that when activated sorbent entered into the sample solution, two valence cations of Pb2+ can create stronger electrostatic interaction by the anionic structure of Fe3O4/MIL-96(Al), compared with the one valence cations (i.e., Na+). Thus, the energy released from the interaction of Pb2+ ions with the surface of sorbent was higher than the required energy for ejection of the one valence Na+ from the surface of sorbent. As a result, the total enthalpy of process was negative and adsorption reaction was exothermic. Also, the inert, low labile and hard acid cations in the sample solution can rapidly create stable interaction by the active O cites on the surface of the sorbent in high temperatures. It leads to their interference for sorption of Pb2+. Therefore, temperature controlling process is a good tool for enhancing selectivity of the sorption process. According to the literature,60 the ΔG° of chemical adsorption is usually −400 to −80 kJ mol−1, and the change of free energy for physical adsorption is usually −20−0 kJ mol−1. The obtained value for ΔG° indicated that sorption process is a physical type, which confirmed by electrostatic interaction60 between Pb2+ cations and anionic structure of Fe3O4/MIL-96(Al) in the sorption step. This value indicated that Pb2+ cations did not create stable coordination bond with the O cites of Fe3O4/MIL-96(Al). The absolute values of ΔG° are gradually decreased with increasing of the temperature, displaying the spontaneous nature of the adsorption process at the low temperatures. Negative values of ΔG° showed that the adsorption process was a spontaneous procedure at all the studied temperatures. Regularly, in the solid−liquid adsorption procedures, adsorption of solute onto the adsorbent is an entropy reductive procedure, whereas desorption of solvent from the surface of sorbent is an entropy growing process.33 It can deduced from the negative value of entropy (ΔS° < 0), that desorption process of the solvent molecules does not occur due to the adsorption procedure of Pb2+ on Fe3O4/MIL-96(Al). On the other hand, water molecules that interacted by Lewis acidic sites of MIL-96(Al) can act as active sites for interaction with Pb2+ ions. In addition, aggregation of sorbent particles during the adsorption of Pb2+ on Fe3O4/MIL-96(Al) is other parameter that can cause decreasing in total entropy of sorption process. Therefore, both parameters, sorption of Pb2+ ions and aggregation of the sorbent particles effect on the reduction of total entropy of the sorption process.

Table 3. Adsorption Capacities of Different Adsorbents for Pb2+ Adsorbent CDpoly-MNPs Fe3O4@SiO2-NH2 MNPs MKC Fe3O4@SA-Zr polymer beads RB-MMNPs γ-MPTMS-SCMNPs cell-EDTA material 75-PMS MnO2 modified biochar CTS/PAM gel Fe3O4/MIL-96(Al)

Conditions

Max adsorption capacity (mg/g)

ref.

pH 5.5 pH 6.0

64.50 100.0

61 62

4.5 < pH pH 5.0

86.1 333.33

63 64

3.0 < pH < 5.5 pH 6 pH 7 pH 5.8 pH 5

79.3 70.4 227.3 421.94 268.0

65 66 67 68 69

pH 5 pH 6.9

138.41 301.5 < Qmax

70 This work

Qmax > 301 mg g−1 for Pb2+. It indicates that Fe3O4/MIL96(Al) has high potential for use in Pb2+ ion uptake from water. In addition to the pH of natural waters is nearly neutral, treatment of water in neutral pH is economic and environmental friendly. There are few works for high efficient removal of Pb2+ ions at neutral pH. Furthermore, it is much easier to isolate the adsorbent from solution using a magnet after adsorption. Application of the Method for Real Sample Analysis. The extraction method was applied for the extraction of Pb2+ ions in seawater sample at optimum conditions, After extraction, the sorbent was eluted by 2.5 mL of HNO3 (0.2 M). The concentration of Pb2+ ions in initial sample was calculated, and the results are presented in Table 4. The Table 4. Extraction of Pb2+ Ions in Seawater Sample Sample

Cinitial ± SD (ng mL−1)

Cadded (ng mL−1)

Cfound ± SD (ng mL−1)

Relative recovery (%)

seawater

4.15 ± 0.32

20 40

23.82 ± 0.93 44.26 ± 1.47

98.35 100.27

efficiency of the method to extraction of Pb2+ ions in seawater sample was confirmed via the analysis of the sample spiked with the known amount of Pb2+ ions. Good conformity was obtained between the added and measured Pb2+ ions contents. The relative recoveries were satisfactory (98.35 and 100.27 %) that certify the validity of the method for the extraction of Pb2+ ions in seawater samples. 3183

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering



CONCLUSION For preparation of cation exchange metal−organic trap, magnetic composite of MIL-96(Al) was synthesized and characterized. It was shown that the Fe3O4/MIL-96(Al) had high potential for release of protons. ζ-Potential measurement indicated the anionic surface of the sorbent in natural aqueous solutions. The porous magnetic framework composite showed high potential for preparation of magnetic cation exchange metal−organic trap, due to high empty cavities among the particles for trapping of the bulky compounds, and indicated high capacity for removal of Pb2+ ions. Thermodynamic parameters demonstrated that the electrostatic interaction between the anionic structures of the adsorbent with hydrated Pb2+ ions. Also, water molecules that interacted by Lewis acidic sites of MIL-96(Al) can act as active sites for interaction with Pb2+ ions. Lewis acid sites of Fe3O4/MIL-96(Al) have an important role in aggregation and producing metal−organic trap during the adsorption process. Comparison of results of the Pb2+ adsorption on Fe3O4/MIL-96(Al) from water samples and other adsorbents showed the higher adsorption capacities of the sorbent than some other sorbents, whereas no pH treatment was needed for natural waters analysis. Because of the potential features of Fe3O4/MIL-96(Al), such as high porosity, high available active Lewis acid sits, negative surface charge and good magnetism properties, it can be used as an original basic material for the facile synthesis of a wide range of porous magnetic adsorbents to fast and easy extraction of different compounds.





lack-of-fit test for response surface quadratic model for percentage extraction (PDF)

AUTHOR INFORMATION

Corresponding Author

*Ali Mehdinia. Tel: +982166944873. Fax: +982166944869. Email: [email protected]. ORCID

Ali Mehdinia: 0000-0001-6080-0681 Author Contributions

These authors contributed equally to this work. Notes

The authors declare no competing financial interest.



REFERENCES

(1) Zhao, X.; Bu, X.; Wu, T.; Zheng, S.-T.; Wang, L.; Feng, P. Selective anion exchange with nanogated isoreticular positive metalorganic frameworks. Nat. Commun. 2013, 4, 2344. (2) Park, J. H.; Choi, K. M.; Jeon, H. J.; Choi, Y. J.; Kang, J. K. In-situ observation for growth of hierarchical metal-organic frameworks and their self-sequestering mechanism for gas storage. Sci. Rep. 2015, 5, 12045. (3) Peng, Y.; Gong, T.; Zhang, K.; Lin, X.; Liu, Y.; Jiang, J.; Cui, Y. Engineering chiral porous metal-organic frameworks for enantioselective adsorption and separation. Nat. Commun. 2014, 5, 4406. (4) Peng, L.; Zhang, J.; Xue, Z.; Han, B.; Sang, X.; Liu, C.; Yang, G. Highly mesoporous metal−organic framework assembled in a switchable solvent. Nat. Commun. 2014, 5, 4465. (5) Zheng, H.; Zhang, Y.; Liu, L.; Wan, W.; Guo, P.; Nyström, A. M.; Zou, X. One-pot Synthesis of Metal−Organic Frameworks with Encapsulated Target Molecules and Their Applications for Controlled Drug Delivery. J. Am. Chem. Soc. 2016, 138 (3), 962−968. (6) An, J.; Shade, C. M.; Chengelis-Czegan, D. A.; Petoud, S.; Rosi, N. L. Zinc-adeninate metal− organic framework for aqueous encapsulation and sensitization of near-infrared and visible emitting lanthanide cations. J. Am. Chem. Soc. 2011, 133 (5), 1220−1223. (7) Shimizu, G. K.; Taylor, J. M.; Kim, S. Proton conduction with metal-organic frameworks. Science 2013, 341 (6144), 354−355. (8) Cook, T. R.; Zheng, Y.-R.; Stang, P. J. Metal−organic frameworks and self-assembled supramolecular coordination complexes: comparing and contrasting the design, synthesis, and functionality of metal− organic materials. Chem. Rev. 2013, 113 (1), 734−777. (9) Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal−organic frameworks for separations. Chem. Rev. 2012, 112 (2), 869−932. (10) Zhu, Q.-L.; Xu, Q. Metal−organic framework composites. Chem. Soc. Rev. 2014, 43 (16), 5468−5512. (11) Zhu, Q.-L.; Li, J.; Xu, Q. Immobilizing metal nanoparticles to metal−organic frameworks with size and location control for optimizing catalytic performance. J. Am. Chem. Soc. 2013, 135 (28), 10210−10213. (12) Sugikawa, K.; Nagata, S.; Furukawa, Y.; Kokado, K.; Sada, K. Stable and functional gold nanorod composites with a metal−organic framework crystalline shell. Chem. Mater. 2013, 25 (13), 2565−2570. (13) deKrafft, K. E.; Wang, C.; Lin, W. Metal-Organic Framework Templated Synthesis of Fe2O3/TiO2 Nanocomposite for Hydrogen Production. Adv. Mater. 2012, 24 (15), 2014−2018. (14) Yu, Z.; Zhang, C.; Zheng, Z.; Hu, L.; Li, X.; Yang, Z.; Ma, C.; Zeng, G. Enhancing phosphate adsorption capacity of SDS-based magnetite by surface modification of citric acid. Appl. Surf. Sci. 2017, 403, 413−425. (15) Zhang, G.; Hou, S.; Zhang, H.; Zeng, W.; Yan, F.; Li, C. C.; Duan, H. High Performance and Ultra-Stable Lithium-Ion Batteries Based on MOF-Derived ZnO@ ZnO Quantum Dots/C Core−Shell Nanorod Arrays on a Carbon Cloth Anode. Adv. Mater. 2015, 27 (14), 2400−2405.

ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.7b03301. Figure S1 shows the schematic of Fe3O4/MIL-96(Al) activation and Pb2+ ions sorption; Figure S2 shows EDS element composition analysis of Fe3O4/MIL-96(Al) composite; Figure S3 shows XRD pattern of Fe3O4/ MIL-96(Al) after activation by NaOH solution; Figure S4 shows nitrogen adsorption−desorption isotherms of the synthesized Fe3O4/MIL-96(Al); Figure S5 shows FT-IR spectra of Fe3O4, synthesized MIL-96(Al) and Fe3O4/MIL-96(Al); Figure S6 shows the sorbent ability in trapping precipitate of Pb2+ and schematic of separation and aggregation mechanism; Figure S7 shows the EDS element composition analysis of the Fe3O4/MIL-96(Al) after adsorption of Pb2+; Figure S8 shows FT-IR for phosphate functionalized Fe3O4/MIL96(Al); Figure S9 shows the effect of sorption parameters on percentage extraction, and 3D surface graphs and interaction plots; Figure S10 shows the plots of Qe and Qe/Ce vs Ce for Pb2+ adsorption onto Fe3O4/ MIL-96(Al); Figure S11 shows the adsorption isotherm for Pb2+ adsorption onto Fe3O4/MIL-96(Al), experimental and predicted by Temkin model; Figure S12 shows the reusability of Fe3O4/MIL-96(Al) for removing of Pb2+; Figure 13 shows the predicted vs actual data for percentage extraction; Table 1 shows the experimental variables and levels of the central composite design (CCD); Table 2 shows the design matrix and responses for central composite design (CCD) in sorption step; Table 3 shows the analysis of variance (ANOVA) and 3184

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering

capacity with temperature. Microporous Mesoporous Mater. 2010, 129 (1−2), 274−277. (33) Qiu, M.; Chen, C.; Li, W. Rapid controllable synthesis of AlMIL-96 and its adsorption of nitrogenous VOCs. Catal. Today 2015, 258, 132−138. (34) Zhang, N.; Yang, X.; Yu, X.; Jia, Y.; Wang, J.; Kong, L.; Jin, Z.; Sun, B.; Luo, T.; Liu, J. Al-1,3,5-benzenetricarboxylic metal−organic frameworks: A promising adsorbent for defluoridation of water with pH insensitivity and low aluminum residual. Chem. Eng. J. 2014, 252, 220−229. (35) Xia, W.; Zhang, X.; Xu, L.; Wang, Y.; Lin, J.; Zou, R. Facile and economical synthesis of metal-organic framework MIL-100(Al) gels for high efficiency removal of microcystin-LR. RSC Adv. 2013, 3 (27), 11007−11013. (36) Khan, N. A.; Lee, J. S.; Jeon, J.; Jun, C.-H.; Jhung, S. H. Phaseselective synthesis and phase-conversion of porous aluminumbenzenetricarboxylates with microwave irradiation. Microporous Mesoporous Mater. 2012, 152, 235−239. (37) Ahnfeldt, T.; Guillou, N.; Gunzelmann, D.; Margiolaki, I.; Loiseau, T.; Ferey, G.; Senker, J.; Stock, N. [Al4(OH)2(OCH3)4(H2Nbdc)3]x xH2O: a 12-connected porous metal-organic framework with an unprecedented aluminum-containing brick. Angew. Chem., Int. Ed. 2009, 48 (28), 5163−6. (38) Sindoro, M.; Jee, A.-Y.; Granick, S. Shape-selected colloidal MOF crystals for aqueous use. Chem. Commun. 2013, 49 (83), 9576− 9578. (39) Pearson, R. G. Hard and Soft Acids and Bases. J. Am. Chem. Soc. 1963, 85 (22), 3533−3539. (40) Vogel, A. I.; Jeffery, G. Vogel’s textbook of quantitative chemical analysis; Longman Scientific Technical: Harlow, U.K., 1989. (41) Mehdinia, A.; Kayyal, T. B.; Jabbari, A.; Aziz-Zanjani, M. O.; Ziaei, E. Magnetic molecularly imprinted nanoparticles based on grafting polymerization for selective detection of 4-nitrophenol in aqueous samples. J. Chromatogr. A 2013, 1283, 82−88. (42) Knebel, A.; Friebe, S.; Bigall, N. C.; Benzaqui, M.; Serre, C.; Caro, J. R. Comparative Study of MIL-96 (Al) as Continuous Metal− Organic Frameworks Layer and Mixed-Matrix Membrane. ACS Appl. Mater. Interfaces 2016, 8 (11), 7536−7544. (43) Khuri, A. I.; Mukhopadhyay, S. Response surface methodology. Wiley Interdisciplinary Reviews: Comput. Stat. 2010, 2 (2), 128−149. (44) Sohrabi, M. R.; Matbouie, Z.; Asgharinezhad, A. A.; Dehghani, A. Solid phase extraction of Cd (II) and Pb (II) using a magnetic metal-organic framework, and their determination by FAAS. Microchim. Acta 2013, 180 (7−8), 589−597. (45) Ahmad, A. A.; Hameed, B. H.; Ahmad, A. L. Removal of disperse dye from aqueous solution using waste-derived activated carbon: Optimization study. J. Hazard. Mater. 2009, 170 (2−3), 612− 619. (46) Gardiner, W. P.; Gettinby, G. Experimental design techniques in statistical practice: A practical software-based approach. Horwood Publishing Limited: Chichester, U.K., 1998. (47) Barathi, M.; Krishna Kumar, A. S.; Kumar, C. U.; Rajesh, N. Graphene oxide−aluminium oxyhydroxide interaction and its application for the effective adsorption of fluoride. RSC Adv. 2014, 4 (96), 53711−53721. (48) Li, D.; Zhou, J.; Chen, X.; Song, H. Amorphous Fe2O3/ Graphene Composite Nanosheets with Enhanced Electrochemical Performance for Sodium-Ion Battery. ACS Appl. Mater. Interfaces 2016, 8 (45), 30899−30907. (49) O’Neill, H. S. C.; Dollase, W. A. Crystal structures and cation distributions in simple spinels from powder XRD structural refinements: MgCr2O4, ZnCr2O4, Fe3O4 and the temperature dependence of the cation distribution in ZnAl2O4. Phys. Chem. Miner. 1994, 20 (8), 541−555. (50) Zhang, J. L.; Srivastava, R. S.; Misra, R. D. K. Core−Shell Magnetite Nanoparticles Surface Encapsulated with Smart StimuliResponsive Polymer: Synthesis, Characterization, and LCST of Viable Drug-Targeting Delivery System. Langmuir 2007, 23 (11), 6342− 6351.

(16) Aguilera-Sigalat, J.; Bradshaw, D. Synthesis and applications of metal-organic framework−quantum dot (QD@ MOF) composites. Coord. Chem. Rev. 2016, 307, 267−291. (17) Lirio, S.; Liu, W.-L.; Lin, C.-L.; Lin, C.-H.; Huang, H.-Y. Aluminum based metal-organic framework-polymer monolith in solidphase microextraction of penicillins in river water and milk samples. J. Chromatogr. A 2016, 1428, 236−245. (18) Asiabi, M.; Mehdinia, A.; Jabbari, A. Preparation of water stable methyl-modified metal−organic framework-5/polyacrylonitrile composite nanofibers via electrospinning and their application for solidphase extraction of two estrogenic drugs in urine samples. J. Chromatogr. A 2015, 1426, 24−32. (19) Yan, A. X.; Yao, S.; Li, Y. G.; Zhang, Z. M.; Lu, Y.; Chen, W. L.; Wang, E. B. Incorporating polyoxometalates into a porous MOF greatly improves its selective adsorption of cationic dyes. Chem. - Eur. J. 2014, 20 (23), 6927−6933. (20) Zou, F.; Chen, Y.-M.; Liu, K.; Yu, Z.; Liang, W.; Bhaway, S. M.; Gao, M.; Zhu, Y. Metal organic frameworks derived hierarchical hollow NiO/Ni/Graphene composites for lithium and sodium storage. ACS Nano 2016, 10 (1), 377−386. (21) Zhou, J.; Li, X.; Yang, L.; Yan, S.; Wang, M.; Cheng, D.; Chen, Q.; Dong, Y.; Liu, P.; Cai, W.; et al. The Cu-MOF-199/single-walled carbon nanotubes modified electrode for simultaneous determination of hydroquinone and catechol with extended linear ranges and lower detection limits. Anal. Chim. Acta 2015, 899, 57−65. (22) Liang, J.; Liu, J.; Yuan, X.; Dong, H.; Zeng, G.; Wu, H.; Wang, H.; Liu, J.; Hua, S.; Zhang, S.; et al. Facile synthesis of aluminadecorated multi-walled carbon nanotubes for simultaneous adsorption of cadmium ion and trichloroethylene. Chem. Eng. J. 2015, 273, 101− 110. (23) Jia, Y.; Wei, B.; Duan, R.; Zhang, Y.; Wang, B.; Hakeem, A.; Liu, N.; Ou, X.; Xu, S.; Chen, Z.; Lou, X.; Xia, F. Imparting biomolecules to a metal-organic framework material by controlled DNA tetrahedron encapsulation. Sci. Rep. 2015, 4, 5929. (24) Ricco, R.; Malfatti, L.; Takahashi, M.; Hill, A. J.; Falcaro, P. Applications of magnetic metal−organic framework composites. J. Mater. Chem. A 2013, 1 (42), 13033−13045. (25) Na, K.; Choi, K. M.; Yaghi, O. M.; Somorjai, G. A. Metal Nanocrystals Embedded in Single Nanocrystals of MOFs Give Unusual Selectivity as Heterogeneous Catalysts. Nano Lett. 2014, 14 (10), 5979−5983. (26) Lu, G.; Li, S.; Guo, Z.; Farha, O. K.; Hauser, B. G.; Qi, X.; Wang, Y.; Wang, X.; Han, S.; Liu, X.; DuChene, J. S.; Zhang, H.; Zhang, Q.; Chen, X.; Ma, J.; Loo, S. C. J.; Wei, W. D.; Yang, Y.; Hupp, J. T.; Huo, F. Imparting functionality to a metal−organic framework material by controlled nanoparticle encapsulation. Nat. Chem. 2012, 4 (4), 310−316. (27) Hu, Y.; Huang, Z.; Liao, J.; Li, G. Chemical bonding approach for fabrication of hybrid magnetic metal−organic framework-5: high efficient adsorbents for magnetic enrichment of trace analytes. Anal. Chem. 2013, 85 (14), 6885−6893. (28) Summerfield, A.; Cebula, I.; Schröder, M.; Beton, P. H. Nucleation and early stages of layer-by-layer growth of metal organic frameworks on surfaces. J. Phys. Chem. C 2015, 119 (41), 23544− 23551. (29) Giakisikli, G.; Anthemidis, A. N. Magnetic materials as sorbents for metal/metalloid preconcentration and/or separation. A review. Anal. Chim. Acta 2013, 789, 1−16. (30) Loiseau, T.; Lecroq, L.; Volkringer, C.; Marrot, J.; Férey, G.; Haouas, M.; Taulelle, F.; Bourrelly, S.; Llewellyn, P. L.; Latroche, M. MIL-96, a porous aluminum trimesate 3D structure constructed from a hexagonal network of 18-membered rings and μ 3-oxo-centered trinuclear units. J. Am. Chem. Soc. 2006, 128 (31), 10223−10230. (31) Maes, M.; Alaerts, L.; Vermoortele, F.; Ameloot, R.; Couck, S.; Finsy, V.; Denayer, J. F.; De Vos, D. E. Separation of C5-hydrocarbons on microporous materials: complementary performance of MOFs and zeolites. J. Am. Chem. Soc. 2010, 132 (7), 2284−2292. (32) Lee, J. S.; Jhung, S. H. Vapor-phase adsorption of alkylaromatics on aluminum-trimesate MIL-96: An unusual increase of adsorption 3185

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186

Research Article

ACS Sustainable Chemistry & Engineering (51) Yang, X.; Wang, X.; Feng, Y.; Zhang, G.; Wang, T.; Song, W.; Shu, C.; Jiang, L.; Wang, C. Removal of multifold heavy metal contaminations in drinking water by porous magnetic Fe2O3@ AlO(OH) superstructure. J. Mater. Chem. A 2013, 1 (3), 473−477. (52) Liu, D.; Liu, Y.; Dai, F.; Zhao, J.; Yang, K.; Liu, C. Size-and morphology-controllable synthesis of MIL-96 (Al) by hydrolysis and coordination modulation of dual aluminium source and ligand systems. Dalton Trans. 2015, 44 (37), 16421−16429. (53) Li, Q.; Jiang, S.; Ji, S.; Shi, D.; Yan, J.; Huo, Y.; Zhang, Q. Magnetically recyclable Cu-BTC@ SiO2@ Fe3O4 catalysts and their catalytic performance for the Pechmann reaction. Ind. Eng. Chem. Res. 2014, 53 (39), 14948−14955. (54) Liu, H.; Jia, Z.; Ji, S.; Zheng, Y.; Li, M.; Yang, H. Synthesis of TiO2/SiO2@Fe3O4 magnetic microspheres and their properties of photocatalytic degradation dyestuff. Catal. Today 2011, 175 (1), 293− 298. (55) Helm, L.; Merbach, A. E. Water exchange on metal ions: experiments and simulations. Coord. Chem. Rev. 1999, 187 (1), 151− 181. (56) Diebler, H.; Eigen, M.; Ilgenfritz, G.; Maass, G.; Winkler, R. Kinetics and mechanism of reactions of main group metal ions with biological carriers. Pure Appl. Chem. 1969, 20 (1), 93−116. (57) Pehlivan, E.; Altun, T. Ion-exchange of Pb2+, Cu2+, Zn2+, Cd2+, and Ni2+ ions from aqueous solution by Lewatit CNP 80. J. Hazard. Mater. 2007, 140 (1), 299−307. (58) Nalaparaju, A.; Jiang, J. Ion Exchange in Metal−Organic Framework for Water Purification: Insight from Molecular Simulation. J. Phys. Chem. C 2012, 116 (12), 6925−6931. (59) Ma, Q.; Song, T.-Y.; Yuan, P.; Wang, C.; Su, X.-G. QDs-labeled microspheres for the adsorption of rabbit immunoglobulin G and fluoroimmunoassay. Colloids Surf., B 2008, 64 (2), 248−254. (60) Qiu, T.; Zeng, Y.; Ye, C.; Tian, H. Adsorption Thermodynamics and Kinetics ofp-Xylene on Activated Carbon. J. Chem. Eng. Data 2012, 57 (5), 1551−1556. (61) Badruddoza, A. Z.; Shawon, Z. B.; Tay, W. J.; Hidajat, K.; Uddin, M. S. Fe3O4/cyclodextrin polymer nanocomposites for selective heavy metals removal from industrial wastewater. Carbohydr. Polym. 2013, 91 (1), 322−32. (62) Emadi, M.; Shams, E. Preconcentration of Pb2+ by iron oxide/ amino-functionalized silica core−shell magnetic nanoparticles as a novel solid-phase extraction adsorbent and its determination by flame atomic absorption spectrometry. J. Iran. Chem. Soc. 2013, 10 (2), 325− 332. (63) Qin, L.; Yan, L.; Chen, J.; Liu, T.; Yu, H.; Du, B. Enhanced Removal of Pb2+, Cu2+, and Cd2+ by Amino-Functionalized Magnetite/ Kaolin Clay. Ind. Eng. Chem. Res. 2016, 55 (27), 7344−7354. (64) Li, X.; Qi, Y.; Li, Y.; Zhang, Y.; He, X.; Wang, Y. Novel magnetic beads based on sodium alginate gel crosslinked by zirconium(IV) and their effective removal for Pb2+ in aqueous solutions by using a batch and continuous systems. Bioresour. Technol. 2013, 142, 611−9. (65) Madrakian, T.; Afkhami, A.; Ahmadi, M. Simple in situ functionalizing magnetite nanoparticles by reactive blue-19 and their application to the effective removal of Pb2+ ions from water samples. Chemosphere 2013, 90 (2), 542−7. (66) Huang, C.; Hu, B. Silica-coated magnetic nanoparticles modified with γ-mercaptopropyltrimethoxysilane for fast and selective solid phase extraction of trace amounts of Cd, Cu, Hg, and Pb in environmental and biological samples prior to their determination by inductively coupled plasma mass spectrometry. Spectrochim. Acta, Part B 2008, 63 (3), 437−444. (67) d’Halluin, M.; Rull-Barrull, J.; Bretel, G.; Labrugère, C.; Le Grognec, E.; Felpin, F.-X. Chemically modified cellulose filter paper for heavy metal remediation in water. ACS Sustainable Chem. Eng. 2017, 5 (2), 1965−1973. (68) Huang, R.; Wu, M.; Zhang, T.; Li, D.; Tang, P.; Feng, Y. Template-free Synthesis of Large-Pore-Size Porous Magnesium Silicate Hierarchical Nanostructures for High-Efficiency Removal of Heavy Metal Ions. ACS Sustainable Chem. Eng. 2017, 5 (3), 2774− 2780.

(69) Liang, J.; Li, X.; Yu, Z.; Zeng, G.; Luo, Y.; Jiang, L.; Yang, Z.; Qian, Y.; Wu, H. Amorphous MnO2 Modified Biochar Derived from Aerobically Composted Swine Manure for Adsorption of Pb (II) and Cd (II). ACS Sustainable Chem. Eng. 2017, 5 (6), 5049−5058. (70) Ma, J.; Zhou, G.; Chu, L.; Liu, Y.; Liu, C.; Luo, S.; Wei, Y. Efficient Removal of Heavy Metal Ions with An EDTA Functionalized Chitosan/Polyacrylamide Double Network Hydrogel. ACS Sustainable Chem. Eng. 2017, 5 (1), 843−851.

3186

DOI: 10.1021/acssuschemeng.7b03301 ACS Sustainable Chem. Eng. 2018, 6, 3176−3186