Fang-ying
Guo
ab,
Yun-guo
Liu
*ab,
Hui
Wang
ab,
Guang-ming
Zeng
ab,
Xin-jiang
Hu
ab,
Bo-hong
Zheng
c,
Ting-ting
Li
ab,
Xiao-fei
Tan
ab,
Shu-fan
Wang
ab and
Ming-ming
Zhang
ab
aCollege of Environmental Science and Engineering, Hunan University, Changsha 410082, PR China. E-mail: nhxyhj111@163.com; Fax: +86 731 88822829; Tel: +86 731 88649208
bKey Laboratory of Environmental Biology and Pollution Control (Hunan University), Ministry of Education, Changsha 410082, PR China
cSchool of Architecture and Art, Central South University, Changsha 410082, PR China
First published on 12th May 2015
A novel magnetic composite adsorbent was synthesized by grafting 1,2-diaminocyclohexanetetraacetic acid to magnetic graphene oxide (DCTA/E/MGO). The DCTA/E/MGO was employed for removing Cr(VI) from aqueous solution in this study. The composite was characterized by FESEM, TEM, BET, XRD, FT-IR and XPS. The adsorption behaviors of Cr(VI) by DCTA/E/MGO in aqueous solution were systematically investigated. Second order kinetic and Freundlich isotherm models validated the experimental data. The adsorption rate was influenced by both film diffusion and intraparticle diffusion. Thermodynamic parameters revealed that the adsorption reaction was an endothermic and spontaneous process. The novel adsorbent exhibited better Cr(VI) removal efficiency in solutions with low pH. The decontamination of Cr(VI) by DCTA/E/MGO was influenced by ionic strength. These results are important for estimating and optimizing the removal of metal ions by the DCTA/E/MGO composite.
Normally, the adsorption ability of a material is controlled in part by the number of available functional groups. Graphene oxide (GO) prepared by Hummers method4 contains a range of oxygen-containing functional groups on the surface, such as hydroxyl, epoxide, carbonyl and carboxyl groups.5 These groups are available for removing heavy metals from wastewater. In addition, GO is characterized by a large specific surface area and can be readily obtained from cheap natural graphite in large scale.6 From these points, GO is considered as a suitable adsorbent for the removal of pollutants in water. However, GO can be dispersed in aqueous media due to the oxygen-containing functional groups on its surface, therefore it is difficult to be separated from water after the adsorption process. The problem can be solved by loading magnetic nanoparticles to GO. The separation can be achieved by magnetic separation. The magnetic technology combines the advantages of adsorption with the merit of easy separation, but it has been found to have some negative effects on the adsorption capacity of the GO.7,8
In order to improve the adsorption ability and selectivity of GO for metal ions, a great number of GO derivates have been obtained by grafting new chemical substances on the GO backbone, such as sulfanilic acid,4 EDTA,5 chitosan9 and ethylenediamine.10 Meanwhile, 1,2-diaminocyclohexane-N,N,N′,N′-tetraacetic acid (DCTA), acting as a multidentate chelating ligand, can form strong complexes with most metals. These complexes have similar chemical structures to those formed with ethylenediaminetetraacetic acid (EDTA) but have higher stability constants for most metals.11,12 Besides, among various functional groups, the amine group has a relatively high reactivity and can easily react with many chemicals.2,8 Therefore, it is a feasible way to graft DCTA onto magnetic graphene oxide (MGO) through the ethylenediamine.
The main objectives of this work were to: (1) prepare DCTA/E/MGO composite and characterize it by FESEM, TEM, BET, XRD, FT-IR and XPS; (2) study the adsorption mechanism with kinetics, isotherm and thermodynamic models; (3) evaluate the effects of process parameters on Cr(VI) removal, i.e., pH, initial ion concentrations and temperature; (4) investigate the effects of ionic strength on Cr(VI) decontamination.
GO was synthesized via modified Hummers method13 from the natural graphite powder. Briefly, graphite powders were first oxidized by concentrated H2SO4, K2S2O8 and P2O5. Next, the concentrated H2SO4, KMnO4 and NaNO3 were used to oxidize the preoxidized graphite, then 30% H2O2 was added to eliminate the excess MnO4−, and the products were rinsed with HCl (10%) and Milli-Q water. The resulting solution was sonicated for 2 h. The magnetic graphene oxide (MGO) was prepared by coprecipitation method. Fe3+ and Fe2+ (molar ratio 2:
1) were mixed in the GO solution with addition of ammonia solution to form Fe3O4–GO composite.4,14
The DCTA/E/MGO was prepared by reacting DCTA with MGO through ethylenediamine. 0.1 M EDC and 0.1 M NHS solution were added to the DCTA dispersion with continuous stirring for 2 h,15,16 then ethylenediamine was added. The mixed system was stirred continuously for 6 h in a water bath at 80 °C after being added into MGO dispersion.8,15 The resulted product was washed repeatedly with Milli-Q water until pH was about neutral and finally stored at room temperature. The preparation sketch of DCTA/E/MGO is shown in Fig. 1.
![]() | (1) |
The N2 adsorption–desorption isotherms (in Fig. 3a) were recorded to investigate the BET surface areas and pore structure of MGO and DCTA/E/MGO composite. The BET surface area of the DCTA/E/MGO was measured to be 310.41 m2 g−1, which is much higher than that of MGO (114.61 m2 g−1). Nitrogen adsorption–desorption analysis shows all the DCTA/E/MGO and MGO samples with micro- and meso-porous structure. The pore size of DCTA/E/MGO samples mainly distribute at 1.4 nm (Fig. 3b). The introduction of DCTA led to a distinct increase in BET surface area which is beneficial for adsorption.
The XRD patterns of GO, MGO and DCTA/E/MGO are shown in Fig. 3c and d. For GO, a strong peak at 2θ = 10.4° occurs, which is the structure expansion as oxygen-containing groups incorporate between the carbon sheets during the course of strong oxidation. For MGO and DCTA/E/MGO, the intense diffraction peaks at the Bragg angles of 30.09, 35.42, 37.05, 43.05, 53.39, 56.94 and 62.51 were observed clearly. These peaks are consistent with the (220), (311), (222), (400), (422), (511) and (440) facets of the cubic spinel crystal planes of Fe3O4 (JCPDS card no. 19-0629), respectively.17 Compared with the pure GO diffracted signals, there are no diffracted signals for the GO sheets in MGO and DCTA/E/MGO, which is ascribed to that the strong signals of the iron oxides tend to overwhelm the weak carbon peaks.18 Almost similar XRD patterns of MGO and DCTA/E/MGO were observed, which revealed that the synthesized process did not change the crystalline phase of Fe3O4. But, it was not sufficient to exclude the existence of γ-Fe2O3. However, the magnetic properties of γ-Fe2O3 are similar to Fe3O4, so there was no negative impact on the following experiments whether γ-Fe2O3 was contained in the samples or not.19
The FT-IR spectra of MGO and DCTA/E/MGO in the range of 4000–500 cm−1 are shown in Fig. 3e. The FTIR spectra of MGO and DCTA/E/MGO show the presence of O–H at 3449 cm−1. In the FT-IR spectrum of MGO, the peak at 1730 cm−1 corresponding to CO of carboxyl group on the GO shifts to 1643 cm−1 may be due to the formation of –COO− after coating with Fe3O4.4 The stretching band of Fe–O peak appears at around 580 cm−1. While in the FT-IR spectrum of DCTA/E/MGO, the bands at 1636 and 1616 cm−1 correspond to the characteristic C
O stretching vibration of –NHCO– (amide I) and the N–H bending of –NH2, respectively.8,20 The peaks at 1035 and 1384 cm−1 correspond to C–O–C stretching vibration and C–OH stretching, respectively.17 After the chemical grafting, the bands at 1636 and 1616 cm−1 appeared, demonstrating that DCTA and ethylenediamine were successfully grafted on MGO.
The chemical state of element in MGO and DCTA/E/MGO was further investigated by XPS. The wide scan XPS spectrum (Fig. 3f) of DCTA/E/MGO shows photoelectron lines at binding energies of about 285, 400, 530, and 711 eV which are attributed to C1s, N1s, O1s, and Fe2p, respectively.21 The XPS survey of DCTA/E/MGO indicates significant amount of N1s comparing to that of MGO (Fig. 3f), which is originated from the grafted ethylenediamine and DCTA. The elemental analysis illustrates a considerable increase in O/C atomic ratio in the DCTA/E/MGO (0.44) comparing to that of the MGO (0.31), which may be attributed to the high O/C atomic ratio of the introduced DCTA. In addition, five different peaks (Fig. 3g) centered at 284.6 eV, 286.3 eV, 286.9 eV, 288.2 eV and 289.0 eV are observed, corresponding to C–C, C–O, C–O–C, CO and O–C
O groups of MGO, respectively.22 The C1s spectrum of DCTA/E/MGO (Fig. 3h) can be curve-fitted into five peak components with binding energies of about 284.6, 285.6, 286.5, 287.5 and 288.9 eV, which attribute to the carbon atoms in the forms of C–C, C–N, C–O, HNC
O and O–C
O species, respectively.8,23,24 Additional C–N and HNC
O species functions are observed. Thus, it can be concluded that DCTA has been grafted successfully to the MGO surface.
![]() | (2) |
![]() | (3) |
The kinetic parameters calculated from the two models are listed in Table 1. From Table 1, it is noticed that the R2 values of pseudo-second-order model (0.999, 0.998, and 0.998) are higher than those of pseudo-first-order model (0.894, 0.911, and 0.955). In addition, the calculated qe values of the pseudo-second-order model agree with the experimental data better than those of the pseudo-first-order model. Based on these data, it can be concluded that the Cr(VI) uptake process complies with the second-order type kinetic reaction, which indicates that the rate-limiting step may be due to the chemical adsorption, high specific surface area and the absence of internal diffusion resistance.8,26,27 It is also notable that the constant rate k2 decreases with the increase of initial Cr(VI) concentration, which may be connected with the longer time required to reach the equilibrium state.28
Kinetic parameter | 10 mg L−1 | 20 mg L−1 | 40 mg L−1 | |
---|---|---|---|---|
Pseudo-first-order model | q e,exp = 66.34 | q e,exp = 70.49 | q e,exp = 76.71 | |
k 1 (min−1) | 2.46 × 10−3 | 2.40 × 10−3 | 2.58 × 10−3 | |
q e,1 (mg g−1) | 25.95 | 30.12 | 36.94 | |
R 2 | 0.894 | 0.911 | 0.955 | |
Pseudo-second-order model | k 2 (g mg−1 min−1) | 5.92 × 10−4 | 4.68 × 10−4 | 3.55 × 10−4 |
q e,2 (mg g−1) | 66.53 | 70.72 | 77.28 | |
R 2 | 0.999 | 0.998 | 0.998 |
To highlight the influence of diffusion on the adsorption mechanism of Cr(VI), the kinetic data were analyzed by applying the intraparticle diffusion model, which can be described as follows:29
qt = Kdt0.5 + I | (4) |
If the plot qt versus t0.5 is linear and passes through the origin, the intraparticle diffusion is the only rate-limiting step; if the plot presents multilinearity, the adsorption process is controlled by two or more steps.28 As can be seen from Fig. 4d, there are three processes controlling the adsorption rate: (i) the first sharper portions may be attributed to the film diffusion; (ii) the second linear portions are the gradual adsorption stages, where the intra-particle diffusion is rate-controlling step; and (iii) the third portions are final equilibrium stages where intra-particle diffusion starts to slow down due to low adsorbate concentration in aqueous solution as well as less number of available adsorption sites.31 Therefore, the overall rate-limiting step involves both film diffusion and intraparticle diffusion.
![]() | (5) |
qe = KFCe1/n | (6) |
The Langmuir and Freundlich adsorption isotherms obtained using the nonlinear method are shown in Fig. 5, and the related parameters of the two models are listed in Table 2. From the correlation coefficients (R2) and the fitting curves, the experimental data are fitted better by the Freundlich model than by the Langmuir model within the studied temperature range. The better fitting of the Freundlich isotherm indicates that the surface of the DCTA/E/MGO is likely to be heterogeneous. Moreover, all the values of Freundlich constant n (Table 2) in this study are within the beneficial adsorption range, which indicates that the DCTA/E/MGO composite can be used as an effective adsorbent.
T (K) | Langmuir model | Freundlich model | ||||
---|---|---|---|---|---|---|
q max (mg g−1) | K L (L mg−1) | R 2 | K F (L mg−1) | n | R 2 | |
288.15 | 77.75 | 0.163 | 0.900 | 29.78 | 4.66 | 0.950 |
303.15 | 83.66 | 0.267 | 0.861 | 40.33 | 5.84 | 0.953 |
318.15 | 92.27 | 0.401 | 0.892 | 53.07 | 7.57 | 0.933 |
The thermodynamic parameters provide in-depth information about internal energy changes that are associated with adsorption. The standard free-energy change (ΔG°), the standard enthalpy change (ΔH°), and the standard entropy change (ΔS°) are calculated from the temperature-dependent adsorption isotherms to evaluate the feasibility and exothermic nature of the adsorption process. These parameters can be calculated using the following equations:
ΔG° = −RT![]() ![]() | (7) |
![]() | (8) |
Temperature (K) | ln![]() |
ΔG° (kJ mol−1) | ΔH° (kJ mol−1) | ΔS° (J K−1 mol−1) | R 2 |
---|---|---|---|---|---|
288.15 | 0.511 | −1.224 | 7.65 | 30.86 | 0.978 |
303.15 | 0.696 | −1.754 | |||
318.15 | 0.811 | −2.145 |
The ΔG° values are negative, which indicated the spontaneity of the adsorption process. Values of ΔG° decreased slightly with the increase of temperature, which revealed the improvement of the adsorption by increasing the temperature. The positive ΔH° value suggested the endothermic nature of adsorption, which agreed well with the result that the adsorption of Cr(VI) increased along with the increase of temperature (Fig. 4). In addition, the positive ΔS° indicated that the degrees of freedom increased at the solid–liquid interface during the adsorption process.3
![]() | ||
Fig. 6 (a) pH profile of Cr(VI) sorption on DCTA/E/MGO and MGO; (b) effect of ionic strength on Cr(VI) sorption (CCr(VI) initial = 20 mg L−1, T = 30 °C, t = 24 h). |
In aqueous solutions, Cr(VI) exists in different ionic forms such as chromate (CrO42−), dichromate (Cr2O72−) and hydrogen chromate (HCrO4−), depending on the solution pH and the Cr(VI) concentration.1 In the pH range of 1.0–6.0, HCrO4− is the predominant Cr(VI) species. As pH increases, the predominant species is CrO42−.3 At low pH, the large number of protons can easily coordinate with the functional groups on the material surface, which makes the material surface more positive. Thus, higher adsorption capacity at low pH can be explained by the strong electrostatic attraction between the positively charged adsorbent surface and the negatively charged chromate ions. Besides, lower adsorption capacity of Cr(VI) in high pH environment may be due to the dual competition of the anions (CrO42− and OH−) adsorbed on the surface of the adsorbent.40
The qmax value of Cr(VI) adsorption on DCTA/E/MGO is about 80 mg g−1, which is obviously higher than that of MGO (48 mg g−1) and other adsorbents, such as ethylenediamine functionalized Fe3O4 (61.35 mg g−1),27 magnetic cyclodextrin–chitosan/graphene oxide (67.66 mg g−1)17 and cyclodextrin/ethylenediamine/magnetic graphene oxide (68.41 mg g−1).8 It can be seen that the DCTA/E/MGO has higher sorption capacity. DCTA and ethylenediamine have two types of reactive functional groups, carboxyl groups and amino groups, which act as chelation sites and increase the adsorption capacity.
![]() ![]() | (9) |
![]() ![]() | (10) |
![]() ![]() | (11) |
![]() ![]() | (12) |
This journal is © The Royal Society of Chemistry 2015 |