Fabrication of Anion-Exchange Polymer Layered Graphene

Dec 5, 2016 - ... of Science and Technology of China, No. 1088 Xueyuan Road, Xili, Nanshan District, Shenzhen 518055, China. ACS Sustainable Chem...
0 downloads 0 Views 4MB Size
Research Article pubs.acs.org/journal/ascecg

Fabrication of Anion-Exchange Polymer Layered Graphene− Melamine Electrodes for Membrane Capacitive Deionization Xiaoyu Gu,† Yonghong Deng,*,‡ and Chaoyang Wang*,† †

Downloaded via NEW MEXICO STATE UNIV on July 5, 2018 at 23:12:57 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Research Institute of Materials Science, South China University of Technology, No. 381 Wushan Road, Tianhe District, Guangzhou 510640, China ‡ Department of Materials Science & Engineering, South University of Science and Technology of China, No. 1088 Xueyuan Road, Xili, Nanshan District, Shenzhen 518055, China S Supporting Information *

ABSTRACT: A novel nitrogen-doped reduced graphene sponge composite (NRGS) is fabricated by using melamine sponge to restrain the aggregation of graphene sheets during reduction. The anion-exchange polymer layered graphene composites (A-NRGS) are prepared by coating the surface of the NRGS electrode with cross-linked poly(vinyl alcohol) with quaternization modification (C-qPVA). With the help of a melamine sponge to suppress the agglomerate of graphene sheets, the NRGS exhibits a unique three-dimensional (3D) interconnected porous structure with abundant nitrogen doping of 5.2%. Its specific surface area is up to 241 m2/g. In addition, the enhanced wettability of A-NRGS composites favors the diffusion of ion from the electrolyte to electrode. Therefore, A-NRGS composites have excellent electrochemical capacity (184 F/g). The membrane capacitive deionization (MCDI) performance for A-NRGS electrode (11.3 mg/g) is higher than that of pristine reduced graphene oxide (RGO) (6.2 mg/g) and NRGS (8.6 mg/g) electrodes. All the results demonstrate that A-NRGS composites can be a promising candidate for CDI and other electrochemical applications. KEYWORDS: Graphene, N-Doped, Anion-exchange, Membrane capacitive deionization, Melamine



INTRODUCTION Fresh water supply has been an impending challenge in modern civilization, as water pollution and contamination become increasingly serious due to the rapid development of modern industry.1,2 Desalination of seawater is an effective way to produce safe drinking water in order to solve the water crisis. Capacitive deionization (CDI) is an electrosorption water treatment based on the formation of the electrical double layers (EDL), which has drawn extensive attention in the water treatment industry because of its energy saving, nonpolluting, and environmental benignity.1,3 At the electrolyte−electrode interface, ionic species with opposite charges are electrosorbed to perforated electrodes under direct electric fields.4 Membrane CDI (MCDI) is a specific CDI technology that introduces ionexchange membranes into the CDI electrodes, which can effectively minimize the desorption due to the electrostatic interaction between oppositely charged ions in the solution.5 Many researches have demonstrated that the electrosorptive capacity (qe) of MCDI is superior to that of CDI without the impact of co-ions effect, which also means that MCDI is more energy efficient.6,7 However, the electrical resistivity might increase significantly because of the existence of the interfaces between the membranes and electrodes, which would influence the diffusion of ions and limit the total ion adsorption amount.8 To overcome this problem, Choi’s group reported a layered carbon electrode with ion-exchange resins for MCDI application. The resistance between the membranes and carbon © 2016 American Chemical Society

electrodes decreased remarkably. The method demonstrated that qe of this membrane-modified carbon electrode was improved 19%.9 With regards to this working principle, the electrodes made up of porous carbon materials with high surface area and appropriate pore size distribution would be an ideal material in the CDI applications. Therefore, porous carbon materials,10−17 carbon nanotube composites,18−25 and graphene,26−29 were intensively investigated. Among the carbon-based species, graphene with outstanding electrical, mechanical, optical, and photovoltaic properties was proven to have excellent CDI performance.26−29 However, the aggregation of graphene during the reduction process of graphene oxide (GO) would generate low conductivity and specific surface area.30−32 To date, different strategies were applied to restrain the aggregation of graphene and improve the CDI performance.33−36 Yan and Sun’s group reported a novel sponge (polyurethane, PU)-templated graphene-based CDI electrode with high surface area.37 Our group has fabricated 3D graphene/metal oxide hybrids,38 graphene/conducting polymers composites,39 3D macroporous graphene architectures,40 and mesoporous graphene sheets,41 to be used as CDI electrodes. Received: July 19, 2016 Revised: November 22, 2016 Published: December 5, 2016 325

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

Figure 1. Schematic diagram of the fabrication of NRGS electrodes for CDI applications. Fabrication of NRGS and A-NRGS Electrodes. NRGS composite (80 wt %), acetylene black (10 wt %), and polytetrafluoroethylene (PTFE, 10 wt %) were ground for 2−3 h. The NRGS electrode was obtained by casting and pressing the grinding mixtures onto a nickel foam plate. In this work, A-NRGS electrode was prepared by casting cross-linked quaternized PVA (C-qPVA) solution onto a NRGS electrode, which provided anion-exchange functional groups and acted as an anion-exchange membrane. Therefore, it is an asymmetric CDI system in which the NRGS electrode has no cationexchange membrane. qPVA was synthesized and cross-linked by glutaraldehyde (GA) according to a previously reported method.44 The detailed experimental procedures can be found in Support Information (SI). The degree of quaternization (Dq) of qPVA and the degree of cross-linking (Dc) of C-qPVA can be calculated according to Equations S1 and S2. Electrochemical Measurements. To investigate the electrochemical performance of the NRGS and A-NRGS electrode, galvanostatic charge/discharge (GC) tests, cyclic voltammetry (CV) analysis, and electrochemical impedance spectroscopy (EIS) were carried out in NaCl aqueous solutions (1 M). The pH value of aqueous electrolyte is neutral. All electrochemical experiments were measured using a three-electrode cell system and more detailed information on CV, GC, and EIS tests can be found in the Supporting Information. The three-electrode method was usually used in the electrochemical tests about a CDI study. The reference electrode has a very stable potential, which can guarantee the stability of potential in the working electrode. It may be appropriate to use the three-electrode cell system to prevent polarization. Electrosorptive Capacity Measurement. The CDI experiments of A-NRGS were performed in a continuous recycle system as described in our pervious works.39,40 The system consists of a CDI cell, a constant flow pump (Longer BT100-1L), a potentiostat (RS1302DQ), a conductivity meter (DDS 307) and a water tank, as shown in Figure S1. The neutral NaCl solution of 150 mL was pumped into the CDI cell by the constant flow pump and the flow rate was maintained at 10 mL/min. The concentration and conductivity calibration curve of NaCl was shown in Figure S2. The electrosorptive capacity (qe, mg/g) and salt removal efficiency (η) of the electrodes were obtained with eqs 1 and 2, respectively.45,46

In this work, we have prepared nitrogen-doped reduced graphene sponge composite (NRGS) based on poly(melamineformaldehyde) (PMF) sponge-template, the surface of which is covered by quaternized poly(vinyl alcohol) (PVA) for membrane capacitive deionization. A unique graphene macroporous structure throughout the entire melamine sponge is well-preserved after the annealing process, and its electrochemical capacity is enhanced due to the contribution of nitrogen doping. More importantly, the obtained graphene composite exhibits excellent electrosorption capacity of 11.3 mg/g in MCDI application. The detailed preparation process of the NRGS electrode for CDI applications is illustrated in Figure 1.



EXPERIMENTAL SECTION

Materials. PVA, (2,3-epoxypropyl) trimethylammonium chloride (EPTMAC) with purity ≥95%, and graphite powder, were obtained from J&K Chemical (China). PMF sponges were provided by Guangzhou GreenCARE Co. Ltd. (China). All other chemicals were of analytical grade and used as received. Water with resistivity ≥18.0 MΩ·cm was used in all experiments. Preparation of NRGS Composites. GO was produced from graphite according to our previous work.41 GO aqueous dispersion (2 mg/mL) was formed by ultrasonication for 2 h. PMF sponge was soaked into the dispersion to absorb GO sheets. To make GO distribute well in the PMF sponge, the soaking-extruding process was repeated more than five times.42 The sponge then was dried in vacuum for 12 h to remove water. Different contents of GO sheets in the sponges can be tuned by increasing the number of the adsorption− drying cycle.43 The GO contents can be calculated by the weight difference of the sponges before and after GO adsorption. Finally, NRGS composite was obtained after the GO-sponge was annealed from 30 to 800 °C with a heating rate of 3 °C/min under argon atmosphere. The samples are denoted by GO-S-n and NRGS-n before and after annealing, respectively, where n represents the number of the adsorption−drying cycles. Unless specifically noted, GO-S refers to GO-S-3 and NRGS refers to NRGS-3, respectively. Pristine reduced graphene oxide (RGO) was also fabricated independently without the PMF sponge template. 326

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

Figure 2. (a−d) Photographs of PMF sponges with different times of the dipping−drying process. From left to right: PMF, GO-S-1, GO-S-2, and GO-S-3 composites. SEM photos of (e) GO-S-3 and (f,g) NRGS-3 composites with different magnifications. (h) SEM image of A-NRGS-3 electrode. Contact angle images of water droplets on (i) NRGS-3 and (j) A-NRGS-3 electrodes coated with C-qPVA polymer.

qe =

(C0 − Ce)V m

of GO into RGO after annealing, as displayed in Figure 2f. Different from polyurethane (PU)-templated graphene composites with a complete pyrolysis, NRGS exhibits a macroporous structure with a well-preserved PMF sponge framework without serious deformation by the carbonization treatment.37 The surface of the sponge is covered by flexible graphene sheets from Figure 2g. The well-preserved 3D PMF sponge network would suppress the agglomeration of graphene, leading to faster ion diffusion and an effective double-layer capacitance. Herein, A-NRGS electrodes are successfully employed as electrodes for membrane capacitive deionization by the introduction of positively charged quaternized poly(vinyl alcohol) groups. The synthesis chemical reaction equation for cross-linked qPVA (C-qPVA) is shown in Scheme S1. According to the Dq calculated by Equation S1, samples qPVA-3 have the largest Dq value as shown in Table S1. It is because KOH plays a role in the quaternization of PVA by opening the epoxy bonding. However, KOH crystals would be generated on the surfaces of the qPVA membrane due to the existence of excessive KOH.44 The hydrophilicity of PVA increases by introducing quaternary ammonium groups, and PVA may seriously swell in water. Therefore, GA is used to cross-link the qPVA membrane to improve its stability in the aqueous solution. Table S2 shows the Dc of C-qPVA, and the weight ratio of qPVA remains at 4 wt % in all conditions. We can find that sample C-qPVA-1 has the largest Dc value when the weight ratio of GA is 0.4 wt %. However, hydrophilicity of the C-qPVA membranes with excess GA content would decrease due to the consumption of quaternary ammonium groups. It is found that the C-qPVA-2 membrane with a Dc value of 4.97 does not dissolve after being immersed into a water bath for 24 h. This suggests that the cross-linking of

(1)

where C0 (mg/mL) and Ce (mg/mL) are the initial and final NaCl concentrations, respectively. V (mL) is the total volume of the NaCl solution, and m (g) is the total mass of the electrodes.

η(%) =

C0 − Ce × 100 C0

(2)

Characterization. The morphology of the obtained materials was observed by scanning electron microscopy (SEM, Zeiss EVO18), and the operating voltage is 10 kV. Wide-angle X-ray diffraction (XRD) tests were measured on an X’pert PRO diffractometer with a Cu Kα radiation (wavelength, 0.154 nm). Fourier-transform infrared spectroscopy (FT-IR) was applied on a spectrometer (NICOLET 6700). The specific surface area and pore size distribution were calculated from the N2 adsorption−desorption isotherms (NovaWin2, Quantachrome). X-ray photoelectron spectroscopy (XPS) was examined by a theta probe spectroscope (ESCA Axis Ultra DLD).



RESULTS AND DISCUSSION Morphology of NRGS. To minimize the restacking effects inside graphene sheets during preparation, unique graphene macroporous structure throughout the entire melamine sponge was well-preserved. Figure 2a−d shows the photographs of PMF and GO-absorbed sponges with different dipping and drying processes. We can see that PMF sponge shows a general trend of color from white to dark while the size and shape of GO-absorbed sponges maintain well after three dipping processes. From the SEM image of GO-S in Figure 2e, GO sheets are uniformly distributed in the gap of the PMF sponge. The sponge-shaped graphene block is obtained by the carbonization of the PMF sponge and the thermal reduction 327

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

Figure 3. (a) TG curves of PMF, GO, and GO-S. (b) XRD patterns of RGO, GO-S-3, and NRGS. (c) XPS spectra of the samples. (d) N 1s peaks of XPS spectrum of NRGS-3.

Figure 4. N2 adsorption isotherms, pore size, and distribution (inset) of (a) RGO, (b) NRGS-3, and (c) GO-S-3.

contact area of the solvent.47 Therefore, the A-NRGS electrode would have a higher CDI performance.39 As shown in Figure 3a, the residual weight of PMF sponge is 16.04% due to the unique chemical structure compared to PU sponge. We can see that the TG curves of GO-S-n composites are similar to those of GO and the PMF sponge, and the residual weight of Gr-S-n composites increase with the increased number of dipping process. According to the loading calculation, the GO loading in GO-S-n composites is 118.7, 141.4, 147.1%, respectively. It also can be confirmed by the SEM images of GO-S-n composites in Figure S3a-c, which show that the amount of GO sheets anchored in PMF network obviously increases with increasing times of the dipping processes. From Figure S 3d-S 3f, graphene is uniformly

qPVA and GA has occurred, and the C-qPVA-2 membrane with a 4.97 Dc value is strong enough to be an ion-exchange membrane for MCDI performance. As demonstrated in Figure 2h, the coating layer shows high adhesion to the NRGS electrode. The C-qPVA membrane is uniform and the membrane thickness is approximate 20 μm. It is known that the surface wettability always affect the CDI performance. As depicted in Figure 2i,j, the NRGS electrode has a contact angle of 130°, which is much larger than that of the A-NRGS electrode (35°). The enhanced wettability of the A-NRGS electrode can be ascribed to the increasing hydrophilicity of C-qPVA with introducing quaternary ammonium groups, which can increase the electrolyte diffusion rate and the 328

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

Figure 5. (a) Comparative CV curves of A-NRGS-n (n = 1, 2, 3) and RGO electrodes at 10 mV/s. (b) Comparative CV curves of A-NRGS-3 and NRGS-3 at 10 mV/s. (c) GC curves of A-NRGS-3 electrodes at different current densities. (d) GC curves of RGO and A-NRGS electrodes with a 1 A/g current density. (e) The iR drop as a function of discharge current density for the A-NRGS-3, NRGS-3, and RGO electrodes. (f) Nyquist profiles of the A-NRGS-n (n = 1, 2, 3) and RGO electrodes. The inset is the plots in the high frequency region.

distribution of RGO, NRGS-3, and GO-S-3. The isotherms of the samples have a representative type-IV profile with a H3-type hysteresis loop, indicating the mesoporous structure of the tested composites.52 The specific surface area of NRGS-3 and GO-S-3 is 241 and 154 m2/g, respectively, which is much larger than that of RGO (120 m2/g). Detailed results about the porous structure are listed in Table S4. The PMF-sponge template is a versatile approach to fabricate graphene electrodes with large specific surface area by preventing the agglomeration of graphene sheets during reduction. Electrochemical Performance. First, CV tests are carried out because they are strongly sensitive to the CDI properties of the fabricated graphene electrodes. From Figure 5a, the current response and CV enclosed area of all A-NRGS-n electrodes are larger than those of the RGO electrode, demonstrating that ANRGS-n electrodes have larger specific capacitance and better electrosorption performance. Figure S5 shows CV curves of ANRGS-n, NRGS-n (n = 1, 2, 3) and RGO electrodes. The potential window is from −0.5 to 0.5 V. CV curves of A-NRGSn electrodes remain almost rectangular and no Faradaic reaction is observed with the scan rate varying from 5 to 50 mV/s, which suggests that the CV behavior belongs to the ideal capacitive electrical double layer.53 Compared with the NRGS3 electrode, the A-NRGS-3 with a positively charged ionexchange membrane displays a more symmetry rectangular CV curve, as shown in Figure 5b. The improved electrochemical performance with highly reversible capacity is attributed to the following reasons: (1) The macroporous structures and large accessible surface area increase the specific capacitance with the help of PMF sponge template, which means that more ions would be eliminated by A-NRGS electrodes at the same direct voltage; (2) A-NRGS electrodes have high wettability and hydrophilicity by introducing C-qPVA polymer on the surface of the NRGS electrode, which can favor the electrode/ electrolyte interactions and facilitate the transport of electrons and ions; (3) Since the electronegativity of C is weaker than that of N, the doping of graphene sheets with N heteroatoms can significantly improve the capacitive performance of A-

distributed in the PMF sponge network and porous NRGS network is maintained well with a slight shrinkage due to calcination. As depicted in Figure 3b, GO-S-3 composites exhibit (001) peak at 2θ = 10.3° and a broad peak at 22.5° for the PMF sponge.48 Both of the pristine RGO and NRGS-n composites show a similar XRD pattern of (002) peak at 2θ = 25°, indicating that the embedding of the PMF sponge has no significant influence on the hierarchical structure of NRGS composites and most of the oxygen-containing groups have been removed. It can be also proven from the FT-IR observation in Figure S4, which shows that peaks located at 800 cm−1 corresponding to the presence of triazine-based groups in NRGS composites disappear.49 Therefore, it is proven that the graphene-based NRGS composites are generated after the carbonization of the PMF sponge. XPS measurements are carried out to analyze the surface chemistry of the samples to confirm the effective nitrogen doping into graphene after pyrolysis. The XPS wide spectra of all samples are evaluated by deconvolution of the C 1s, O 1s, and N 1s core elements level. Obviously in Figure 3c, only the signal of oxygen and carbon atoms appears while no N signal can be observed from GO and RGO in the XPS spectra. Compared with GO-S-3, a more strong N 1s peak of NRGS is centered at 398.5 eV (5.2 atom % N), while the oxygen content greatly reduces (9.6 atom % O). More detailed information about atomic percentage can be found in Table S3. Three peaks at 398.5, 399.1, and 401.3 eV can be derived from the N 1s spectrum of the NRGS-3 (Figure 3d), which are attributed to pyridine nitrogen, graphitic nitrogen, and pyrrole nitrogen, respectively.49,50 From the results of XRD, FT-IR, and XPS, it is concluded that nitrogen-doped graphene has been successfully prepared by annealing, and the reduction of GO and carbonization of the PMF sponge was also performed. Pore size and distribution and surface area are important for electrode composites.51 The porous structure of the samples is also investigated by a N2 adsorption−desorption measurement. Figure 4 shows the N2 adsorption isotherms, pore size, and 329

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

Figure 6. (a) CDI performance of the RGO, NRGS, and A-NRGS electrodes at 1.6 V. (b) Regenerated electrosorptive capacity of A-NRGS electrode. (c) The desalination efficiency and electrosorption capacity of the A-NRGS electrode as a function of the conductivity of the initial NaCl solutions.

CDI Performance. To investigate the electrosorption behaviors and selectivity of the anion-exchange layer of ANRGS electrodes, the MCDI test is measured in the 150 mL NaCl solution under 1.6 V. The initial conductivity is 300 μS/ cm. To compensate for the resistance of the CDI system, the potential difference in this test exceeds the threshold of 1.23 V.58 No gas bubbles were observed at the voltage of 1.6 V and no hydrolysis happens in all cases. So, all CDI properties are measured at 1.6 V.40 The electrochemical test and CDI test have some differences. The potential difference of 1.6 V in electrosorption experiments exceeds the threshold of 1.23 V to compensate for the circuit system resistance, while −0.5 to 0.5 V is an appropriate potential window for graphene electrodes to measure the electrochemical performance. From Figure 6a, the conductivity would decline when the voltage is applied to all graphene-based electrodes, indicating a fast ionic adsorption. The conductivity decreases slowly until no change is observed any further with increasing time, which indicates that saturation is achieved because of the mutual repulsion among the ions adsorbed in the electrodes.45 Obviously, the A-NRGS electrodes have larger conductivity change compared to pristine RGO and NRGS electrodes. The electrosorptive capacity of A-NRGS is calculated to be 11.3 mg/g from eq 2, which is much higher than that of 8.6 and 6.2 mg/g for NRGS and RGO, respectively. As discussed above and in other reports, the A-NRGS electrode having a higher adsorption amount can be attributed to the macroporous structure, abundant nitrogen doping, and the selective shielding of the membrane. In comparison with RGO, A-NRGS have protection from the PMF sponge during the annealing process, which leads to successful nitrogen doping at the same time. This provides fine surface accessibility which facilitates the A-NRGS electrodes to easily adsorb the ions from the salt solution. On the other hand, the co-ion effect is effectively eliminated in MCDI because cations can be blocked in the bulk solution with the existence of the anion-exchange membrane.5 Therefore, compared to the CDI performance of

NRGS electrodes due to the pseudofaradic reactions involving the surface active groups.54,55 The GC curves of the A-NRGS-3 electrode at various current densities are shown in Figure 5c. We can find that the GC curves of A-NRGS-3 electrodes are linear and symmetric and the potential drop (iR drop) is low. It indicates that the ANRGS-3 electrodes have excellent stability in the charge/ discharge process. The specific capacitance of A-NRGS-3, ANRGS-2, A-NRGS-1 electrodes calculated from the GC curve in Figure 5d by Equation S3 is 184, 155, 138 F/g, respectively, while the specific capacitance of RGO is 76 F/g. Under the same electrochemical conditions, the longer GC time of the ANRGS electrodes means a higher ion storage capacity, which is attributed to the enlarged specific surface area and abundant nitrogen doping mentioned above. As depicted in Figure 5e, the iR drop is lower at a small current density. It increases gradually with increasing current densities. Furthermore, the iR drop of the A-NRGS-3 electrode is much lower than that of the NRGS3 and pristine RGO electrodes at the same current density. This means that the overall resistance reduces significantly.56 The above results are also consistent with the Nyquist profiles of EIS spectra, which can be seen in Figure 5f. The slope for the A-NRGS-n electrodes at low frequency region is obviously higher than that of the RGO electrodes, which indicates a better capacitive property and faster ionic diffusion into the macroporous graphene composites for the A-NRGS-n electrodes. The equivalent series resistance of the A-NRGS-3 electrode (Rs, the intersection of Nyquist profiles with the real axis) can be calculated to be 0.03 Ω, much smaller than that of RGO (5.46 Ω), which suggests good conductivity. In addition, the negligible semicircle of the A-NRGS-3 electrode indicates low interfacial electron-transfer resistance due to the high wettability of the C-qPVA treated graphene electrode.57 All the results discussed above indicate that the A-NRGS-n electrodes with excellent electrochemical performance are more appropriate as CDI electrode materials in comparison with the pristine RGO and NRGS electrodes. 330

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

sponge and C-qPVA polymer endowed A-NRGS with excellent electrochemical performance because of high surface surface area, abundant nitrogen doping, increased hydrophilicity, and reduced co-ion effect. Compared to RGO and NRGS electrodes, the obtained A-NRGS composite exhibits excellent electrosorption capacity of 11.3 mg/g in MCDI application, as well as good cycling stability and regeneration performance. The results demonstrate that A-NRGS composites not only are promising candidates in CDI application, but also can be applied in a variety of applications such as energy storage and supercapacitors.

RGO and NRGS electrode, the salt removal efficiency of ANRGS can be significantly enhanced. The workable electrodes with a good reversibility are important to the long life for the actual CDI application. To investigate the stability and reversibility of the A-NRGS electrodes, consecutive electrosorption and regeneration experiments are carried out by turning off the voltage of the cell. The regeneration cycles of the E-Gr-Fe3O4 electrodes with the absence of an external potential can be found in Figure S6. From Figure 6b, the regenerated electrosorptive capacities of ANRGS are 11.3, 10.7, 10.3, 9.7, and 9.3 mg/g for the five cycles, respectively. There is little change in the MCDI performance, so the good repeatability of A-NRGS electosorption can be realized with without secondary pollution and with free driving energy. The negligible decline during the operation cycle could have been caused by the incomplete desorption: a small portion of adsorbed ions remains inside the porous electrode in the desorption process, leading to higher ionic density at the surface of the electrodes.59 On the other hand, active sites for ion adsorption also decrease due to consumption of the Faradaic reactions between functional groups on the A-NRGS electrode.15 So the A-NRGS electrodes have a better desalination performance. The number of adsorbed ions would gradually decease because of the limited ion adsorption capacity of the electrode.60 The influence of the initial salt concentrations on the electrosorption capacities of the A-NRGS electrode is shown in Figure 6c. The electrosorption capacities are 8.07, 11.3, 12.8, 14.1, 14.8 mg/g and the desalination efficiencies are 23.4, 11, 7.4, 5.1, and 4.3% with an increase in the initial NaCl concentrations from 100, 300, 500, and 800 to 1000 μs/cm, respectively. Apparently, the total electrosorption capacity is enhanced and the desalination efficiency is reduced with increasing ion concentration.46 This occurs because of the overlapping influence of the EDLs on the decrease of micropores for the ionic resistance at the high ion concentration; therefore, electrosorptive capacities increase.61 On the other hand, the excessive ions would not be electroadsorbed by the vacancy at the high ion concentration, which is inversely proportional to the removal efficiency.62 In fact, ANGRS composites prepared in this paper have a relatively higher capacitance and electrosorption capacity than other graphene-based electrodes which are used in CDI performance.63−65 However, limited to the existing testing conditions, we cannot obtain the corresponding current response during CDI experiments. Therefore, we cannot calculate the current efficiency. In addition, the electrosorptive capacity of A-NRGS is much higher than that of NRGS and RGO, which further convincingly confirm the selective shielding of the membrane since cations can be blocked in the bulk solution with the existence of the anion-exchange membrane. In fact, A-NGRS composites prepared in this paper have a relatively higher capacitance and electrosorption capacity than other graphenebased electrodes which are used in CDI performance.63−65



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.6b01685. Schematic of the CDI device, SEM photos of GO-S, FTIR spectra of GO, GO-S, NRGS, and PMF, CV curves of A-NRGS and RGO electrodes (PDF)



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected] (C.W.) *E-mail: [email protected] (Y.D.). ORCID

Chaoyang Wang: 0000-0002-7270-5451 Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS C.W. thanks NSFC for financial support (21274046 and 21474032). REFERENCES

(1) Anderson, M. A.; Cudero, A. L.; Palma, J. Capacitive Deionization as an Electrochemical Means of Saving Energy and Delivering Clean Water. Comparison to Present Desalination Practices: Will It Compete? Electrochim. Acta 2010, 55, 3845−3856. (2) Subramani, A.; Badruzzaman, M.; Oppenheimer, J.; Jacangelo, J. G. Energy Minimization Strategies and Renewable Energy Utilization for Desalination: A Review. Water Res. 2011, 45, 1907−1920. (3) Yeh, C. L.; Hsi, H. C.; Li, K. C.; Hou, C. H. Improved Performance in Capacitive Deionization of Activated Carbon Electrodes with a Tunable Mesopore and Micropore Ratio. Desalination 2015, 367, 60−68. (4) Oren, Y. Capacitive Deionization (Cdi) for Desalination and Water Treatment - Past. Desalination 2008, 228, 10−29. (5) Zhao, Y. J.; Wang, Y.; Wang, R. G.; Wu, Y. F.; Xu, S. C.; Wang, J. X. Performance Comparison and Energy Consumption Analysis of Capacitive Deionization and Membrane Capacitive Deionization Processes. Desalination 2013, 324, 127−133. (6) Li, H. B.; Zou, L. Ion-Exchange Membrane Capacitive Deionization: A New Strategy for Brackish Water Desalination. Desalination 2011, 275, 62−66. (7) El-Deen, A. G.; Choi, J.-H.; Kim, C. S.; Khalil, K. A.; Almajid, A. A.; Barakat, N. A. M. TiO2 Nanorod-Intercalated Reduced Graphene Oxide as High Performance Electrode Material for Membrane Capacitive Deionization. Desalination 2015, 361, 53−64. (8) Nie, C. Y.; Pan, L. K.; Liu, Y.; Li, H. B.; Chen, T. Q.; Lu, T.; et al. Electrophoretic Deposition of Carbon Nanotubes−Polyacrylic Acid Composite Film Electrode for Capacitive Deionization. Electrochim. Acta 2012, 66, 106−109.



CONCLUSIONS Graphene composite NRGS was prepared on the basis of the PMF sponge-template, which could restrain the aggregation of graphene sheets during reduction, and lead to successful nitrogen doping after pyrolysis. Then a novel nitrogen-doped A-NRGS composite was fabricated, the surface of which was coated by C-qPVA polymer as anion-exchange membrane for MCDI application. The synergistic effect from a combination of 331

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering

Mixed Matrix Membrane for Co2 Capture. ACS Sustainable Chem. Eng. 2015, 3, 1819−1829. (29) Wang, H.; Zhang, D. S.; Yan, T. T.; Wen, X. R.; Shi, L. Y.; Zhang, J. P. Graphene Prepared Via a Novel Pyridine−Thermal Strategy for Capacitive Deionization. J. Mater. Chem. 2012, 22, 23745−23748. (30) Wang, H.; Shi, L. Y.; Yan, T. T.; Zhang, J. P.; Zhong, Q. D.; Zhang, D. S. Design of Graphene-Coated Hollow Mesoporous Carbon Spheres as High Performance Electrodes for Capacitive Deionization. J. Mater. Chem. A 2014, 2, 4739−4750. (31) Li, H. B.; Zaviska, F.; Liang, S.; Li, J.; He, L. J.; Yang, H. Y. A High Charge Efficiency Electrode by Self-Assembling Sulphonated Reduced Graphene Oxide onto Carbon Fibre: Towards Enhanced Capacitive Deionization. J. Mater. Chem. A 2014, 2, 3484−3491. (32) Zhang, D. S.; Wen, X. R.; Shi, L. Y.; Yan, T. T.; Zhang, J. P. Enhanced Capacitive Deionization of Graphene/Mesoporous Carbon Composites. Nanoscale 2012, 4, 5440−5446. (33) Lei, H.; Yan, T. T.; Wang, H.; Shi, L. Y.; Zhang, J. P.; Zhang, D. S. Graphene-Like Carbon Nanosheets Prepared by a Fe-Catalyzed Glucose-Blowing Method for Capacitive Deionization. J. Mater. Chem. A 2015, 3, 5934−5941. (34) Fan, W.; Zhang, C.; Tjiu, W. W.; Pramoda, K. P.; He, C. B.; Liu, T. X. Graphene-Wrapped Polyaniline Hollow Spheres as Novel Hybrid Electrode Materials for Supercapacitor Applications. ACS Appl. Mater. Interfaces 2013, 5, 3382−3391. (35) Jiao, T. F.; Zhao, H.; Zhou, J. X.; Zhang, Q. R.; Luo, X. N.; Hu, J.; et al. Self-Assembly Reduced Graphene Oxide Nanosheet Hydrogel Fabrication by Anchorage of Chitosan/Silver and Its Potential Efficient Application toward Dye Degradation for Wastewater Treatments. ACS Sustainable Chem. Eng. 2015, 3, 3130−3139. (36) Wen, X. R.; Zhang, D. S.; Yan, T. T.; Zhang, J. P.; Shi, L. Y. Three-Dimensional Graphene-Based Hierarchically Porous Carbon Composites Prepared by a Dual-Template Strategy for Capacitive Deionization. J. Mater. Chem. A 2013, 1, 12334−12344. (37) Yang, Z. Y.; Jin, L. J.; Lu, G. Q.; Xiao, Q. Q.; Zhang, Y. X.; Jing, L.; et al. Sponge-Templated Preparation of High Surface Area Graphene with Ultrahigh Capacitive Deionization Performance. Adv. Funct. Mater. 2014, 24, 3917−3925. (38) Gu, X. Y.; Yang, Y.; Hu, Y.; Hu, M.; Wang, C. Y. Fabrication of Graphene-Based Xerogels for Removal of Heavy Metal Ions and Capacitive Deionization. ACS Sustainable Chem. Eng. 2015, 3, 1056− 1065. (39) Gu, X. Y.; Yang, Y.; Hu, Y.; Hu, M.; Huang, J.; Wang, C. Y. Facile Fabrication of Graphene−Polypyrrole−Mn Composites as High-Performance Electrodes for Capacitive Deionization. J. Mater. Chem. A 2015, 3, 5866−5874. (40) Gu, X. Y.; Yang, Y.; Hu, Y.; Hu, M.; Huang, J.; Wang, C. Y. Nitrogen-Doped Graphene Composites as Efficient Electrodes with Enhanced Capacitive Deionization Performance. RSC Adv. 2014, 4, 63189−63199. (41) Gu, X.; Hu, M.; Du, Z.; Huang, J.; Wang, C. Fabrication of Mesoporous Graphene Electrodes with Enhanced Capacitive Deionization. Electrochim. Acta 2015, 182, 183−191. (42) Yang, Y.; Liu, Z. J.; Huang, J.; Wang, C. Y. Multifunctional, Robust Sponges by a Simple Adsorption−Combustion Method. J. Mater. Chem. A 2015, 3, 5875−5881. (43) Nguyen, D. D.; Tai, N.-H.; Lee, S.-B.; Kuo, W.-S. Superhydrophobic and Superoleophilic Properties of Graphene-Based Sponges Fabricated Using a Facile Dip Coating Method. Energy Environ. Sci. 2012, 5, 7908−7912. (44) Zhang, Q. G.; Liu, Q. L.; Zhu, A. M.; Xiong, Y.; Ren, L. Pervaporation Performance of Quaternized Poly(Vinyl Alcohol) and Its Crosslinked Membranes for the Dehydration of Ethanol. J. Membr. Sci. 2009, 335, 68−75. (45) Jia, B. P.; Zou, L. Graphene Nanosheets Reduced by a MultiStep Process as High-Performance Electrode Material for Capacitive Deionisation. Carbon 2012, 50, 2315−2321. (46) Wimalasiri, Y.; Mossad, M.; Zou, L. Thermodynamics and Kinetics of Adsorption of Ammonium Ions by Graphene Laminate

(9) Kim, J. S.; Choi, J. H. Fabrication and Characterization of a Carbon Electrode Coated with Cation-Exchange Polymer for the Membrane Capacitive Deionization Applications. J. Membr. Sci. 2010, 355, 85−90. (10) Porada, S.; Weinstein, L.; Dash, R.; van der Wal, A.; Bryjak, M.; Gogotsi, Y.; et al. Water Desalination Using Capacitive Deionization with Microporous Carbon Electrodes. ACS Appl. Mater. Interfaces 2012, 4, 1194−1199. (11) Huang, Z.; Lu, L.; Cai, Z. X.; Ren, Z. Y. J. Individual and Competitive Removal of Heavy Metals Using Capacitive Deionization. J. Hazard. Mater. 2016, 302, 323−331. (12) Niu, R.; Li, H. B.; Ma, Y. L.; He, L. J.; Li, J. An Insight into the Improved Capacitive Deionization Performance of Activated Carbon Treated by Sulfuric Acid. Electrochim. Acta 2015, 176, 755−762. (13) Shi, K. Y.; Zhitomirsky, I. Influence of Chemical Structure of Dyes on Capacitive Dye Removal from Solutions. Electrochim. Acta 2015, 174, 588−595. (14) Haro, M.; Rasines, G.; Macias, C.; Ania, C. O. Stability of a Carbon Gel Electrode When Used for the Electro-Assisted Removal of Ions from Brackish Water. Carbon 2011, 49, 3723−3730. (15) Xu, P.; Drewes, J. E.; Heil, D.; Wang, G. Treatment of Brackish Produced Water Using Carbon Aerogel-Based Capacitive Deionization Technology. Water Res. 2008, 42, 2605−2617. (16) Ma, T. Y.; Liu, L.; Yuan, Z. Y. Direct Synthesis of Ordered Mesoporous Carbons. Chem. Soc. Rev. 2013, 42, 3977−4003. (17) Li, L. X.; Zou, L.; Song, H. H.; Morris, G. Ordered Mesoporous Carbons Synthesized by a Modified Sol−Gel Process for Electrosorptive Removal of Sodium Chloride. Carbon 2009, 47, 775−781. (18) Shi, K. Y.; Zhitomirsky, I. Supercapacitor Devices for Energy Storage and Capacitive Dye Removal from Aqueous Solutions. RSC Adv. 2015, 5, 320−327. (19) Zhi, X.; Zhang, H. B.; Liao, Y. F.; Hu, Q. H.; Gui, C. X.; Yu, Z. Z. Electrically Conductive Polycarbonate/Carbon Nanotube Composites Toughened with Micron-Scale Voids. Carbon 2015, 82, 195−204. (20) Li, H. B.; Liang, S.; Gao, M. M.; Li, G. L.; Li, J.; He, L. J. The Study of Capacitive Deionization Behavior of a Carbon Nanotube Electrode from the Perspective of Charge Efficiency. Water Sci. Technol. 2015, 71, 131−136. (21) Estevez, L.; Dua, R.; Bhandari, N.; Ramanujapuram, A.; Wang, P.; Giannelis, E. P. A Facile Approach for the Synthesis of Monolithic Hierarchical Porous Carbons − High Performance Materials for Amine Based Co2 Capture and Supercapacitor Electrode. Energy Environ. Sci. 2013, 6, 1785−1790. (22) Hatzell, K. B.; Fan, L.; Beidaghi, M.; Boota, M.; Pomerantseva, E.; Kumbur, E. C.; et al. Composite Manganese Oxide Percolating Networks as a Suspension Electrode for an Asymmetric Flow Capacitor. ACS Appl. Mater. Interfaces 2014, 6, 8886−8893. (23) Peng, Z.; Zhang, D. S.; Yan, T. T.; Zhang, J. P.; Shi, L. Y. ThreeDimensional Micro/Mesoporous Carbon Composites with Carbon Nanotube Networks for Capacitive Deionization. Appl. Surf. Sci. 2013, 282, 965−973. (24) Shi, K. Y.; Ren, M.; Zhitomirsky, I. Activated Carbon-Coated Carbon Nanotubes for Energy Storage in Supercapacitors and Capacitive Water Purification. ACS Sustainable Chem. Eng. 2014, 2, 1289−1298. (25) Lu, M.; Y, L. J.; Chen, J.; Wang, S. P.; Yang, J. M. Functionalized Graphene/Activated Carbon Composite Electrodes for Asymmetric Capacitive Deionization. Acta Phys-Chim Sin 2014, 30, 2263−2271. (26) Liu, M. K.; Tjiu, W. W.; Pan, J. S.; Zhang, C.; Gao, W.; Liu, T. X. One-Step Synthesis of Graphene Nanoribbon-Mno2 Hybrids and Their All-Solid-State Asymmetric Supercapacitors. Nanoscale 2014, 6, 4233−4242. (27) Yang, Y.; Ren, L. l.; Zhang, C.; Huang, S.; Liu, T. X. Facile Fabrication of Functionalized Graphene Sheets (Fgs)/Zno Nanocomposites with Photocatalytic Property. ACS Appl. Mater. Interfaces 2011, 3, 2779−2785. (28) Shen, Y. J.; Wang, H. X.; Liu, J. D.; Zhang, Y. T. Enhanced Performance of a Novel Polyvinyl Amine/Chitosan/Graphene Oxide 332

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333

Research Article

ACS Sustainable Chemistry & Engineering Electrodes in Capacitive Deionization. Desalination 2015, 357, 178− 188. (47) Xiong, Y.; Fang, J.; Zeng, Q. H.; Liu, Q. L. Preparation and Characterization of Cross-Linked Quaternized Poly(Vinyl Alcohol) Membranes for Anion Exchange Membrane Fuel Cells. J. Membr. Sci. 2008, 311, 319−325. (48) Yang, Y.; Tong, Z.; Ngai, T.; Wang, C. Y. Nitrogen-Rich and Fire-Resistant Carbon Aerogels for the Removal of Oil Contaminants from Water. ACS Appl. Mater. Interfaces 2014, 6, 6351−6360. (49) Xiao, N.; Lau, D.; Shi, W. H.; Zhu, J. X.; Dong, X. C.; Hng, H. H.; et al. A Simple Process to Prepare Nitrogen-Modified Few-Layer Graphene for a Supercapacitor Electrode. Carbon 2013, 57, 184−190. (50) Zhao, Z. K.; Li, W. Z.; Dai, Y. T.; Ge, G. F.; Guo, X. W.; Wang, G. R. Carbon Nitride Encapsulated Nanodiamond Hybrid with Improved Catalytic Performance for Clean and Energy-Saving Styrene Production Via Direct Dehydrogenation of Ethylbenzene. ACS Sustainable Chem. Eng. 2015, 3, 3355−3364. (51) Xu, X. T.; Sun, Z.; Chua, D. H. C.; Pan, L. K. Novel Nitrogen Doped Graphene Sponge with Ultrahigh Capacitive Deionization Performance. Sci. Rep. 2015, 5, 11225−11234. (52) Wang, H.; Zhang, D. S.; Yan, T. T.; Wen, X. R.; Zhang, J. P.; Shi, L. Y.; et al. Three-Dimensional Macroporous Graphene Architectures as High Performance Electrodes for Capacitive Deionization. J. Mater. Chem. A 2013, 1, 11778−11789. (53) Hou, C. H.; Liu, N. L.; Hsi, H. C. Highly Porous Activated Carbons from Resource-Recovered Leucaena Leucocephala Wood as Capacitive Deionization Electrodes. Chemosphere 2015, 141, 71−79. (54) Liu, Y.; Xu, X. t.; Wang, M.; Lu, T.; Sun, Z.; Pan, L. K. NitrogenDoped Carbon Nanorods with Excellent Capacitive Deionization Ability. J. Mater. Chem. A 2015, 3, 17304−17311. (55) Rasines, G.; Lavela, P.; Macías, C.; Zafra, M. C.; Tirado, J. L.; Ania, C. O. On the Use of Carbon Black Loaded Nitrogen-Doped Carbon Aerogel for the Electrosorption of Sodium Chloride from Saline Water. Electrochim. Acta 2015, 170, 154−163. (56) Zhang, D. S.; Yan, T. T.; Shi, L. Y.; Peng, Z.; Wen, X. R.; Zhang, J. P. Enhanced Capacitive Deionization Performance of Graphene/ Carbon Nanotube Composites. J. Mater. Chem. 2012, 22, 14696− 14704. (57) Xu, X. Z.; Zhou, J.; Nagaraju, D. H.; Jiang, L.; Marinov, V. R.; Lubineau, G.; et al. Flexible, Highly Graphitized Carbon Aerogels Based on Bacterial Cellulose/Lignin: Catalyst-Free Synthesis and Its Application in Energy Storage Devices. Adv. Funct. Mater. 2015, 25, 3193−3202. (58) Li, H. B.; Zou, L.; Pan, L. K.; Sun, Z. Novel Graphene-Like Electrodes for Capacitive Deionization. Environ. Sci. Technol. 2010, 44, 8692−8697. (59) Wang, G.; Qian, B. Q.; Dong, Q.; Yang, J. Y.; Zhao, Z. B.; Qiu, J. S. Highly Mesoporous Activated Carbon Electrode for Capacitive Deionization. Sep. Purif. Technol. 2013, 103, 216−221. (60) Lee, J. B.; Park, K. K.; Eum, H. M.; Lee, C. W. Desalination of a Thermal Power Plant Wastewater by Membrane Capacitive Deionization. Desalination 2006, 196, 125−134. (61) Jeon, S.; Park, H.; Yeo, J.; Yang, S. C.; Cho, C. H.; Han, M. H.; et al. Desalination Via a New Membrane Capacitive Deionization Process Utilizing Flow-Electrodes. Energy Environ. Sci. 2013, 6, 1471− 1475. (62) Li, H. B.; Liang, S.; Li, J.; He, L. J. The Capacitive Deionization Behaviour of a Carbon Nanotube and Reduced Graphene Oxide Composite. J. Mater. Chem. A 2013, 1, 6335−6341. (63) Yasin, A. S.; Mohamed, H. O.; Mohamed, I. M. A.; Mousa, H. M.; Barakat, N. A. M. Enhanced Desalination Performance of Capacitive Deionization Using Zirconium Oxide NanoparticlesDoped Graphene Oxide As a Novel and Effective Electrode. Sep. Purif. Technol. 2016, 171, 34−43. (64) Wang, G.; Dong, Q.; Wu, T. T.; Zhan, F.; Zhou, M.; Qiu, J. S. Ultrasound-Assisted Preparation of Electrospun Carbon Fiber/ Graphene Electrodes for Capacitive Deionization: Importance and Unique Role of Electrical Conductivity. Carbon 2016, 103, 311−317.

(65) Xu, X. T.; Liu, Y.; Wang, M.; Yang, X. X.; Zhu, C.; Pan, L. K. Design and Fabrication of Mesoporous Graphene via Carbothermal Reaction for Highly Efficient Capacitive Deionization. Electrochim. Acta 2016, 188, 406−413.

333

DOI: 10.1021/acssuschemeng.6b01685 ACS Sustainable Chem. Eng. 2017, 5, 325−333