Yanzhen Guo,
Binbin Chang,
Ting Wen,
Chunmei Zhao,
Hang Yin,
Yannan Zhou,
Yonggang Wang,
Baocheng Yang* and
Shouren Zhang*
Institute of Nanostructured Functional Materials, Huanghe Science and Technology, College, Zhengzhou, Henan 450006, China. E-mail: ybc@hhstu.edu.cn; zhangshouren1984@126.com
First published on 4th February 2016
A facile one-step microwave irradiation method is developed to synthesize a composite of graphene and zinc oxide, which is quick and environmentally friendly. The obtained nanostructured hybrid material was utilized for the fabrication of a supercapacitor electrode and the electrochemical performance was investigated intensively. The electrode with the as-prepared graphene/zinc oxide composite displayed a specific capacitance as high as 201 F g−1 in 1 M Na2SO4 at a current density of 1 A g−1, which is much higher than those of pure microwave reduction graphene oxide (77 F g−1) and zinc oxide (5 F g−1) at the same current density. In addition, the specific capacitance of the MRGO/ZnO electrode presented an outstanding cycling stability with capacitance retention of 93% even after 3000 cycles. This research provides a feasible method for low-cost manufacture of other graphene-based materials with high-performance applications in energy-storage.
Among the different kinds of carbon-based materials, graphene has emerged as the most appealing material for supercapacitor applications in recent years.10,11 Graphene is a one-atom-thick two-dimensional sheet with high conductivity and ultrahigh theoretic surface area (over 2600 m2 g−1) which facilitates electron transfer between the active materials and the charge collector. Unfortunately, the unavoidable agglomeration of graphene dramatically restricts the surface area, the amount of active sites and impedes the effective electron transfer, resulting in an unsatisfactory capacitive property. Studies have shown that the specific capacitances of graphene can only achieve 135 and 99 F g−1 in aqueous and organic electrolytes, respectively,7 far less than the theoretical capacitance of 550 F g−1. In order to reduce the aggregation and thus improve the specific capacitance, many methods have been employed to enhance the supercapacitor performance of graphene-based materials. One way of the methods is the application of a modified synthesis route. A three-dimensional holey graphene framework with a hierarchical porous structure is synthesized by adding H2O2 in the hydrothermal reaction, and the graphene framework electrode delivers a gravimetric capacitance of 298 F g−1.12 A porous carbon with a Brunauer–Emmett–Teller surface area of 3100 m2 g−1 is obtained by using chemical activation of microwave exfoliated graphite oxide, delivering a specific capacitance of 165 F g−1 at a current density of 1.4 A g−1.13 Similarly, graphene-derived carbons obtained via microwave expansion and KOH activation exhibit a specific capacitance of 174 F g−1 in an ionic liquid electrolyte.14 Additionally, 3D graphene sponges are also reported to improve the capacitance of graphene.15–17 Another approach to enhance the supercapacitor performance is using pseudocapacitive materials18 (e.g. RuO2,19 MnO2,20 conductive polymer,21–23 etc.) as additives to graphene which can decrease the aggregation and restacking of graphene sheets,24,25 leading the composite obtain a higher specific capacitance.
Zinc oxide, as a wide-band gap semiconductor, has been used as a potential candidate for supercapacitor application for its eco-friendly nature. However, ZnO usually delivers a relatively low specific capacitance due to the poor electrical conductivity and accessible surface areas. Thus ZnO always act as an additives to carbon materials,26,27 especially for graphene.28,29 Graphene/ZnO composite film synthesized by deposition of ZnO on graphene film via ultrasonic spray pyrolysis exhibits a specific capacitance of 61.7 F g−1, which is higher than that of graphene (38.9 F g−1).29 Li et al. prepared the composite of graphene–ZnO by hydrothermal process, possessing a high capacitance of 156 F g−1 at 5 mV s−1.30 Now there are more and more methods are proposed to prepare graphene/zinc oxide composites, including hydrothermal,30,31 solvothermal,32 thermal decomposition,33 supercritical carbon dioxide,34 etc. However, the synthetic methods are always time-consuming or need intricate manipulation. Besides, the addition of a chemical reductant, such as hydrazine or NaBH4 is always needed for the reduction of graphene oxide, which is potentially hazardous to human. Therefore, an environmental friendly and facile approach for the synthesis of graphene/ZnO is urged to propose.
Microwave irradiation has been used in the exfoliation and reduction of graphene oxide to form graphene,35,36 as well as for the preparation of graphene/carbon nanotubes composites.37,38 Researches have shown that reduction of graphene oxide by microwave irradiation can improve the specific surface area and enhance the supercapacitor performance.13,14 In this work, we report a cost-effective microwave irradiation synthesis method for the production of microwave reduction graphene oxide/zinc oxide (MRGO/ZnO) nanocomposite by just augment a portion of zinc acetate to graphene oxide before the process of microwave irradiation. The as-prepared MRGO/ZnO composite was used as supercapacitor electrode material and exhibited a specific capacitance of 201 F g−1 in 1 M Na2SO4 at a current density of 1 A g−1, which is much higher than those of pure MRGO (77 F g−1) and ZnO (5 F g−1) under identical test condition. Furthermore, the specific capacitance of MRGO/ZnO electrode displayed an outstanding cycling stability with capacitance retention of 93% after 3000 cycles.
:
1 in aqueous solution were grinded in a mortar for about 30 min, until the solvent was fully volatilized. Then the mixture was transferred to a glass vial, placed in a domestic microwave oven. MRGO/ZnO composite was obtained by simple microwave irradiation of the mixture for 5 min at 1000 W.
:
1
:
1. After thorough mixing, the slurry was coated onto the nickel foam (1 × 1 cm2) and then dried at 80 °C for 12 h in a vacuum oven. The mass of active material for MRGO and MRGO/ZnO loading in each working electrode was approximately 2 mg, while 5–6 mg for ZnO. All electrochemical measurements were carried out in 1 M Na2SO4 aqueous electrolyte at room temperature by using a CHI660D electrochemical workstation. The electrochemical impedance spectroscopy (EIS) was performed at open circuit potential over a frequency range from 0.01 to 100
000 Hz. Cyclic voltammetry (CV) curves were performed at a scan rate ranging from 10 to 100 mV s−1 with a potential window from −1 to 0 V. The galvanostatic charge–discharge (GCD) measurements were conducted with a current density of 1 to 5 A g−1 with the same potential window as that for the CV tests.
The specific capacitance values were calculated from galvanostatic charge–discharge curves according to the following equations:
In a three-electrode system:
![]() | (1) |
In a two-electrode system:
![]() | (2) |
The energy density (E, W h kg−1) and power density (P, kW kg−1) of the cell were calculated based on the two-electrode configuration according to the following equations:
![]() | (3) |
![]() | (4) |
:
1, the preparation of graphene/zinc oxide showed a favorable reproducibility and electrochemical performance.
The morphologies of MRGO, ZnO and MRGO/ZnO were examined by SEM and TEM, as shown in Fig. 2. MRGO with wrinkled edges and 3–5 layers was clearly observed from the images of Fig. 2a–c, which was also reported by our previously work.36 ZnO nanoparticles synthesized by hydrothermal method were highly crystallized with a diameter range of 50 to 100 nm (Fig. 2d–f). Fig. 2g shows the digital photos of the mixture of GO and Zn(CH3COO)2·2H2O before and after microwave treatment, the color of the mixture turned brown to black after microwave irradiation for 5 min at 1000 W, indicating the reduction of graphene oxide and decomposition of Zn(CH3COO)2·2H2O. ZnO nanoparticles were uniformly dispersed on the surface of graphene, as seen in Fig. 2h–k. The diameter of ZnO nanoparticles in MRGO/ZnO was about 20 nm, much smaller and more uniform than the ZnO nanoparticles synthesized by hydrothermal method. In the HRTEM image of MRGO/ZnO (Fig. 2l), graphene edge of MRGO was clearly observed. An interplanar spacing of 0.26 nm for the lattice fringes corresponds to that of ZnO (002) crystal plane.
TG measurement was carried out to determine the weight ratio of ZnO to graphene in the composites of MRGO/ZnO, as shown in Fig. 3a. The TG and derivative thermogravimetric (DTG) curves afforded information about the degradation extent of MRGO/ZnO. Only one step at about 535 °C for mass loss appeared, which is due to the removal of MRGO from the composites. The calculations based on the TG and DTG curves indicated that the ZnO content in MRGO/ZnO was about 50% and the decomposition of MRGO was processed at the temperature of 450–610 °C.
![]() | ||
| Fig. 3 (a) TG and DTG curves of MRGO/ZnO. XRD patterns (b), FTIR spectra (c), Raman spectra (d) of ZnO, MRGO and MRGO/ZnO. | ||
Fig. 3b shows the XRD patterns of MRGO, ZnO and MRGO/ZnO. After reduction of GO by microwave irradiation, a broad peak at 26.6° appeared in MRGO, manifesting the successful reduction and exfoliation of GO. The XRD pattern of ZnO consists of six characteristic diffraction peaks at 31.8, 34.4, 36.2, 47.5, 56.6 and 62.8°, corresponding to the (100), (002), (101), (102), (110) and (103) planes of the hexagonal ZnO phase (JCPDS 36-1451), respectively. In the XRD pattern of MRGO/ZnO, there was a broad peak at about 26° besides the characteristic peaks of ZnO, indicating the successful combination of MRGO and ZnO.
The structure of MRGO/ZnO was further supported by the FTIR spectrum, as shown in Fig. 3c. The broad absorption band at ∼3450 cm−1 ascribes to the O–H stretching vibrations of absorbed water on the ZnO surface. The absorption band in the range of 450–460 cm−1 can be assigned to the vibrations of Zn–O,44 which can be observed in the spectra of both ZnO and MRGO/ZnO. No stretching vibrations of O–H at ∼3400 cm−1 are observed in the spectra of MRGO and MRGO/ZnO, indicating the successful reduction of GO by microwave irradiation. Both the absorption peaks of MRGO and the stretching vibration of Zn–O can be observed in the spectrum of MRGO/ZnO, confirming the successful preparation of MRGO/ZnO.
Raman scattering is highly sensitive to the electronic structure and is an important tool for characterization of carbonaceous materials. For the spectrum of ZnO (Fig. 3d), the intense peak at ∼435 cm−1 corresponds to non-polar optical phonon E2, which is the characteristic peak of the hexagonal-crystalline phase ZnO,45,46 in accordance with the XRD results, while the peak of 1150 cm−1 is due to the multiple-phonon scattering processes.47 G band around 1595 cm−1 is related to the phonon vibrations of sp2 hybridized carbons, while D band at 1345 cm−1 is corresponding to the disordered carbon. The intensity ratio ID/IG of MRGO/ZnO (0.95) exhibits a higher value than that of MRGO (0.38) due to the increasing disorder of sp2 contributed by the presence of ZnO in the composite.
The UV-Vis absorption spectra and photoluminescence spectra of as-prepared materials were also collected (Fig. 4). ZnO shows an absorption peak at ∼375 nm, which can also be observed in MRGO/ZnO, while MRGO exhibits no obvious peak in the range of 200–800 nm. PL spectra of the as-synthesized materials were characterized to examine the optical properties. With the excitation wavelength of 375 nm, an emission peak at ∼430 nm was appeared in ZnO, which can also be found in MRGO/ZnO. Two other weak emission peaks at ∼490 and 535 nm ascribed to oxygen-vacancies were clearly observed in ZnO, while indiscoverable in the nanocomposite, implying few defects and highly crystalline of ZnO in MRGO/ZnO composites.34
![]() | ||
| Fig. 4 UV-Vis absorption spectra of ZnO, MRGO and MRGO/ZnO. The inset shows the corresponding photoluminescence spectra. | ||
Fig. 5 presents the N2 adsorption–desorption isotherm and pore size distribution of MRGO, ZnO and MRGO/ZnO. MRGO (Fig. 5a) displays a typical type IV isotherm with a clear H3 hysteresis loop, indicative of the existence of flakiness structure with slit pore. Meanwhile, the large adsorption quantity of nitrogen implies a high surface area of 465.7 m2 g−1 for MRGO, which favors the obtaining of a satisfactory surface area of MRGO/ZnO. As expected, MRGO/ZnO (Fig. 5c) exhibits the type IV isotherms with a H4 hysteresis loop at a relative pressure from 0.4 to 1.0, indicating the generation of porous structure with a wide pore size. More importantly, MRGO/ZnO possesses a prominent surface area of 109.5 m2 g−1, which is much higher than those of pure ZnO (9.2 m2 g−1) and even the reported value of G-ZnO (7.9 m2 g−1).48 The pore size distribution of MRGO/ZnO (the insert Fig. 5c) shows that there are two types of pore existing in the framework structure. The primary pore is typical mesopore with 2.43 nm size, which should be attributed to the interspaces between the layers of graphene sheets, and this result is consistent with pore wide of MRGO (the insert Fig. 5a). Another larger pore is mesopore with a widen pore distribution of 7.6–12.8 nm size, which is probably related to the space between the sheets in ZnO nanoparticles. Thus, the satisfactory surface area and superior porosity of MRGO/ZnO endow the advantage for the efficient electron adsorption and transfer.
![]() | ||
| Fig. 5 Nitrogen adsorption isotherms of MRGO (a), ZnO (b) and MRGO/ZnO (c). The inset graphs are the pore size distribution. | ||
:
1, for that of 1
:
1 and 4
:
1, the composites showed a less uniform morphology and an unfavorable electrochemical performance which is due to the deficient exertion of graphene or ZnO (Fig. S2 and Table S1†). While the weight of graphene oxide was further decreased to 1/2 for that of Zn(CH3COO)2·2H2O, the composites of MRGO/ZnO cannot be gained on account of the lack of microwave absorbents. The results of Fig. 6 reveal that the incorporation of ZnO into MRGO can greatly improve the capacitive performance of the sample, which can be attributed to cases as follows: (1) 2D structure of graphene allows for the uniform distribution of ZnO, which provides larger electrochemically active surface areas for charge transfer than that of ZnO; (2) the existence of ZnO and the wrinkle nature of graphene facilitate the ion diffusion during charge/discharge process; (3) the pseudocapacitance obtained from the reaction between ZnO and Na2SO4 electrolyte, which is mainly result from the intercalation and deintercalation of Na+ from electrolyte into ZnO:28 ZnO + Na+ + e− ↔ ZnONa enlarged the CV loops. In general, the introduction of ZnO may allow comprehensive utilization of the EDL capacitance of graphene and pseudocapacitance of ZnO to produce synergistic effects.
| Materials | Synthetic method | Specific capacitance | Electrolyte | Reference number |
|---|---|---|---|---|
| ZnO–graphene | Zn powder reduction | 192 F g−1 (2 mV s−1) | KOH | 50 |
| Graphene–ZnO film | Ultrasonic spray pyrolysis | 11.3 F g−1 (30 mV s−1) | 1 M KCl | 51 |
| Graphene–ZnO | Ultrasonic spray pyrolysis | 61.7 F g−1 (50 mV s−1) | 1 M KCl | 29 |
| Graphene/ZnO | Microwave-assisted method | 109 F g−1 (5 mV s−1) | 1 M KCl | 52 |
| Graphene/ZnO | Thermal reduction | 62.2 F g−1 (5 mA cm−2) | 1 M KOH | 53 |
| Graphene–ZnO | Hydrothermal reaction | 156 F g−1 (5 mV s−1) | 0.5 M Na2SO4 | 30 |
| Graphene–ZnO | Microwave-assisted reduction | 146 F g−1 (2 mV s−1) | 1 M KCl | 54 |
| Graphene/ZnO | Microwave irradiation | 201 F g−1 (1 A g−1) | 1 M Na2SO4 | Our work |
In order to further understand the electrochemical behavior of the samples, EIS measurements are performed (Fig. 7). In the complex plane, the imaginary component, Z′′ shows the capacitive property while the real component, Z′ displays the ohmic property. In the high-frequency region, the real axis intercept represents the equivalent series resistance (ESR), which is the sum of resistance of the electrolyte, the contact resistance of the interface active material/current collector and the intrinsic resistance of the active material.55 The ESR value of MRGO obtained from the x-intercept is 0.65 Ω, which endows that of MRGO/ZnO (1.16 Ω) lower than ESR of ZnO (1.55 Ω). At low frequency, the imaginary part sharply increases and a nearly vertical line is observed, indicating an ideal EDL capacitive electrode material.56 The Warburg curve of MRGO/ZnO at low frequency is shorter than that of MRGO and ZnO, indicating lower electronic resistance and faster ion diffusion in the sample. In the medium frequency domain, a small semicircle reflecting the electrochemical reaction impedance of the electrode is observed, and a smaller semicircle means smaller charge transfer resistance.57,58 It is apparent that the charge transfer resistance for MRGO (∼0.82 Ω) and MRGO/ZnO (∼1.14 Ω) are much lower than that of ZnO (∼3.41 Ω). The low ESR is crucial for enhancing rate capability or the power density of the electrochemical capacitors since the power density is inversely to ESR.57
![]() | ||
| Fig. 7 Nyquist plots of MRGO, ZnO and MRGO/ZnO, inset is the expanded high-frequency region of the plots. | ||
Furthermore, we studied the durability of the electrode materials using galvanostatic charge–discharge measurement at a current density of 2 A g−1 for 3000 cycles, as shown in Fig. 8. It is clear that the specific capacitance still remains of 165 F g−1 (about 93% of the initial capacitance) after 3000 cycles, which indicates that MRGO/ZnO displays a good cycling and stable energy-storage process as the supercapacitive electrode material. However, the capacitance retention of MRGO and ZnO remains only for 90% and 48%, respectively. The inset of Fig. 8 shows the galvanostatic charge–discharge curve of MRGO/ZnO electrode at a current density of 2 A g−1, all the curves are linear, symmetrical and no obvious IR drop is observed, illustrating a low thermal resistance owing to the well-formed electrode/electrolyte interface.34
![]() | ||
| Fig. 8 Cyclic stability of MRGO, ZnO and MRGO/ZnO at a current density of 2 A g−1 for 3000 cycles. Inset is the charge–discharge profile for the MRGO/ZnO electrode. | ||
A two-electrode test cell can provide more accurate measurement of the materials performance for supercapacitor.49,59 Therefore, the supercapacitor performance of the samples was further investigated by a symmetrical two-electrode system. The CV curves of MRGO/ZnO composite with different scan rates ranging from 10 to 100 mV s−1 are shown in Fig. 9a. CV curves of the composite show rectangular shape even when the scan rate increases up to 100 mV s−1, indicating the ideal capacitive behavior, which is consistent with the results in three-electrode system. Fig. 9b shows the GCD curves of MRGO/ZnO with different current densities. The symmetrical triangular shape of the curves suggests the excellent electrochemical reversibility of the material. The specific capacitance of the composite calculated from the galvanostatic charge–discharge at 1 A g−1 is 125 F g−1, which is lower than the value of 201 F g−1 based on three-electrode configuration. This phenomenon has been observed in the previous reports,59,60 which can be attributed to the distinct electrode polarization effect in different electrodes. The CV curves at a scan rate of 50 mV s−1 for the samples were as well collected, as shown in Fig. 9c. The encircled surface of the CV curve for MRGO/ZnO is much larger than those of MRGO and ZnO, indicating significantly improved capacitive properties of the material, which is also observed in the GCD curves in Fig. 9c. EIS spectra of the electrode materials in a symmetric cell were also provided in Fig. S3,† and the shape of the curves is similar to that in three-electrode system. As two of the most primary parameters to determine the energy storage capability for supercapacitors, energy and power density are estimated and presented in Fig. 9d. MRGO/ZnO delivers energy density of 4.37 and 3.06 W h kg−1 with the corresponding power density of 0.53 and 3.42 kW kg−1 at 1 and 5 A g−1, higher than that of the pure MRGO powder electrode (1.35 W h kg−1 with the power density of 0.57 kW kg−1 at 1 A g−1). Nevertheless, MRGO presents a power density of 5 kW kg−1 at 5 A g−1, higher than that of MRGO/ZnO (3.42 kW kg−1), which is due to the lower ESR of MRGO. Such results clearly demonstrate the preferable cycling stability and high specific capacitance of our materials for high-performance supercapacitors.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra24212f |
| This journal is © The Royal Society of Chemistry 2016 |