Yan-Hong Ding,
Jun Peng*,
Hai-Yang Lu,
Yue Yuan and
Shifa-Ullah Khan
Key Laboratory of Polyoxometalate Science of Ministry of Education, Faculty of Chemistry, Northeast Normal University, Changchun, Jilin 130024, PR China. E-mail: jpeng@nenu.edu.cn
First published on 3rd August 2016
Polyoxometalates (POMs) based inorganic–organic composite materials have attracted considerable interest for energy storage and conversion applications. Modification of substrate materials is an effective method to achieve ideal electrochemical performance. Herein, poly(dimethyl diallyl ammonium chloride) (PDDA) functionalized graphene (PDDA–RGO) was used as the conductive matrix to support tungsten addenda mixed heteropolymolybdate, PMo12−xWxO403− (PMoW), with a facile, straightforward one-pot method. The cationic polyelectrolyte, PDDA, was employed as a linker to adsorb both RGO and POMs through electrostatic interactions. The resulting PMoW–PDDA–RGO composite exhibited a homogeneous honeycomb-like porous structure leading to fast ion transport and short ion diffusion pathways. In a typical two-electrode symmetric system, the supercapacitor exhibited a high rate specific capacitance of 140 F g−1 even at a high current density of 10 A g−1 and an energy density of 6.15 W h kg−1 at a power density of 250 W kg−1. After 1700 cycles at a fast discharge speed of 5 A g−1, the capacitance remained 94.6% of its initial capacitance.
Among various redox-active materials, POMs have been gradually identified as a promising candidate of electrode materials for energy storage systems due to their fascinating molecular and electronic structures, chemical tunability and rich redox properties.11–13 However, unfavorable natures of POMs include poor electronic conductivity and partial degradation in aqueous media.14–16 To overcome these problems, researchers are trying to design hybrid materials consisting of pseudocapacitive POMs and conductive substrates such as conducting organic polymers17,18 and carbon materials.19,20 Such types of hybrid materials may sufficiently utilize the unique chemical reactivity of POMs and the electronic properties of conductive substrates.
To date, the classical Keggin-type POM, 12-molybdophosphoric acid (PMo12), has been doped in the conductive polymer polypyrrole (PPy), and the PMo12/PPy hybrid shows a specific capacitance of 130 F g−1 at scan rate of 1 mV s−1.21 Another example of a hybrid material is PMo12/CNTs with a specific capacitance of 40 F g−1 and an energy density on a level of 1.3 W h kg−1.22 Gomez-Romero and co-workers anchored PMo12 to AC, which provided a specific capacity of 160 F g−1 at a current density of 2 A g−1.23 Subsequently, the same group synthesized a PMo12 and RGO nanomaterial (PMo12/RGO) via hydrothermal reactions with a specific capacitance of 123 F g−1 at a current density of 1 A g−1.24 Moreover, they used another classical Keggin-type POM, 12-phosphotungstic acid (PW12), for preparing a PW12/AC composite by ultrasonic processing with a specific capacitance of 254 F g−1 at 10 mV s−1 in a 3-electrode configuration and an extended voltage range up to 1.6 V in an acidic aqueous electrolyte.25
Moreover, the Keggin-type POMs with dual addenda metal elements, such as Mo/W and Mo/V mixed POMs, have been used in combination with multi walled carbon nanotubes (MWCNTs) for electrochemical capacitors (EC).26–29
In view of these basic research, we designed a new hybrid system in which three functional constituents were included: (1) tungsten addenda mixed heteropolymolybdates PMo12−xWx (denoted as PMoW) that are expected to obtain a more ideal rectangular capacitive profile with a synergistic effect between the Mo and W addenda atoms. (2) Reduced graphene oxide (RGO) that has an excellent conductivity and has shown the greatest potential for the development of electrodes with a high power density.30,31 (3) Poly(dimethyl-diallyl ammonium chloride) (PDDA) that is used to change the surface of RGO from negatively charged to positively charged, thus reinforcing the integration between the PMoW and the RGO.32,33 In this study, we explore a one-step hydrothermal method34,35 to prepare the three-component material, and the PMoW–PDDA–RGO ternary composite shows a honeycomb-like porous structure. Through combination of the rich electrochemical redox properties of PMoW, the effective connection of PDDA and the good electrical conductivity of RGO, the degradation of PMoW is avoided and remarkable supercapacitive performances with a high specific capacitance, good rate capacity and cycle stability have been obtained from the PMoW–PDDA–RGO composite in an acidic aqueous electrolyte (1 M H2SO4).
![]() | ||
Fig. 1 (a) The FT-IR spectra and (b) TGA curves (in air) of PMoW, GO, RGO, and the PMoW–PDDA–RGO composite. |
The content of PMoW in the composite is calculated to be 36.2 wt% on the basis of the TGA results (Fig. 1b). It is noted that the weight loss started at about 200 °C for GO, but at about 400 °C for both RGO and the PMoW–PDDA–RGO composite. This fact also supports the effective removal of oxygen-containing functional groups from GO.
Fig. 2a shows the X-ray diffraction (XRD) patterns obtained from GO, RGO, PMoW and PMoW–PDDA–RGO composite. The typical diffraction peak at about 10° for GO disappears for RGO, which evidences again the effective reduction of GO. Moreover, the typical peaks of the pristine PMoW become very flat for the PMoW–PDDA–RGO composite, which further demonstrates that the PMoW molecules are homogeneously dispersed in the composite.
![]() | ||
Fig. 2 (a) The XRD patterns of GO, RGO, PMoW and PMoW–PDDA–RGO composite. (b) The Raman spectra of GO, RGO and PMoW–PDDA–RGO composite. |
The structural and electronic properties of the PMoW–PDDA–RGO composite are further explicated by Raman spectroscopy. As shown in Fig. 2b, two remarkable peaks located at around 1355 and 1600 cm−1 are observed in each Raman spectra, corresponding to the D band (associated with structural defects) and the G band (for the stretching vibration of sp2 carbon), respectively. The intensity ratios of the two bands (ID/IG) are 0.82 for GO, 0.90 for RGO and 0.98 for PMoW–PDDA–RGO. The value of ID/IG represents the electronic conjugation state of a carbonaceous material, and an increase in the value of ID/IG from GO to PMoW–PDDA–RGO implies that more defects and wrinkles have been generated.
XPS of GO, RGO and the PMoW–PDDA–RGO composite are shown in Fig. S2 (ESI†). In the XPS of the PMoW–PDDA–RGO composite, the signals of C, O, N, Mo and W further prove the successful combination of PMoW and PDDA–RGO. Deconvolutions of the C 1s spectra also reveal the effective reduction of GO (Fig. 3). The relative signals of C–O (286.5 eV) and O–CO (288.8 eV) obtained from RGO and the PMoW–PDDA–RGO composite are obviously weaker than those obtained from GO, which confirm again that most of the epoxide, hydroxyl, and carboxyl functional groups have been removed. These results are consistent with the results of IR, XRD and TGA.
Morphology analysis for the PMoW–PDDA–RGO composite has been done by SEM and TEM measurements, as well as for the PMoW–RGO composite as a comparison. The SEM image of the PMoW–PDDA–RGO composite presents a honeycomb-like interconnected network (Fig. 4a), which provides a larger electrolyte-penetration surface and interpenetrating channels for fast electron and ion transports.37 The TEM image further verifies the homogeneous interconnected mesoporous architecture present in the PMoW–PDDA–RGO composite (Fig. 4c). The HRTEM image obtained from a slice of the PMoW–PDDA–RGO composite shows consecutive lattice fringes of PMoW arrays, which explicates that PMoW molecules are discretely and uniformly dispersed in the PDDA–RGO sheets (inset in Fig. 4c). Fig. S3 (ESI†) shows the N2 adsorption/desorption isotherm and the pore-size distribution of the PMoW–PDDA–RGO composite. The BET surface area of the PMoW–PDDA–RGO composite is 143.7 m2 g−1 and an average pore size of about 2.6 nm is calculated by the Barrett–Joyner–Halenda (BJH) method. The PMoW–RGO composite obviously shows a state of aggregation (Fig. 4b and d). This phenomenon may result from two aspects. On the one hand, the binding force between the PMoW and the RGO components will be weak without the presence of PDDA. On the other hand, RGO sheets possess a tendency to aggregate in a strong acidic solution of POMs under hydrothermal conditions.38
![]() | ||
Fig. 4 SEM images of the (a) PMoW–PDDA–RGO and (b) PMoW–RGO composites. (c) TEM image of the PMoW–PDDA–RGO, (inset is HRTEM analysis of a slice of composite). (d) TEM image of the PMoW–RGO composite. |
PMoVI12−xWxO403− + 2e− + 2H+ ⇄ H2PMoV2MoVI10−xWxO403− | (1) |
H2PMoV2MoVI10−xWxO403− + 2e− + 2H+ ⇄ H4PMoV4MoVI8−xWxO403− | (2) |
H4PMoV4MoVI8−xWxO403− + 2e− + 2H+ ⇄ H6PMoV6MoVI6−xWxO403− | (3) |
The redox reactions of WVI/WV centres, which occur below −0.2 V, out of the voltage window of the studied SCs, are ignored here (see Fig. S4, ESI†).25,29,39 The CV results imply that the PMoW–PDDA–RGO composite electrode combines double-layer capacitance, mainly provided by the RGO, and pseudocapacitance, mainly provided by the PMoW.11
Fig. 5b presents the CV curves of the PMoW–PDDA–RGO composite at various scan rates ranging from 2 to 200 mV s−1. It is observed that with an increase in the scan rates, the current densities are gradually enhanced and the three pairs of redox peaks are still well-defined, indicating a good rate capability. In addition, the peak currents increase linearly with the increased scan rate (Fig. 5b, inset), suggesting that the redox-process is surface-confined.25,32
Fig. 5c gives the GCD curves at a current density of 1 A g−1. The GCD of the PMoW–PDDA–RGO composite is not as smooth as that obtained from RGO, and the ripples just correspond to the redox waves in the CV. Calculated from the discharge curves, the specific capacitances are 223.7 F g−1 for RGO and 279.1 F g−1 for the PMoW–PDDA–RGO composite. The enhancement in the specific capacitance of the PMoW–PDDA–RGO composite can be attributed to a synergistic effect between the unique conductivity of the RGO and the rich electrochemical properties of the PMoW, as well as the effective connection between the RGO and PMoW by PDDA. In addition, factors such as the obtained homogeneous honeycomb-like porous structure leading to fast ion transport and short ion diffusion pathways, uniform dispersion of pseudocapacitive PMoW and inhibition of the degradation of PMoW, are also beneficial for the super capacitance.
To further evaluate the electrochemical rate performances of the as-prepared electrode material, specific capacitances were measured at different current densities (Fig. 5d). The specific capacitance of the PMoW–PDDA–RGO composite electrode was 253.6 F g−1 at a current density of 2 A g−1 and retained 224 F g−1 at a current density of 20 A g−1. While the RGO electrode displayed a laggard behaviour with 216.4 F g−1 at a current density of 2 A g−1 and 186 F g−1 at a current density of 20 A g−1.
In order to realize a practical application for the PMoW–PDDA–RGO composite in SCs, a symmetric SC device was assembled in a 2-electrode system. The GCD curves were recorded at a current density of 10 A g−1 in the potential range of 0–1 V in a 1 M H2SO4 aqueous solution (Fig. 6a). Fig. 6a clearly shows a longer discharge time for the PMoW–PDDA–RGO, implying a larger specific capacitance than that of RGO. Fig. 6b shows the GCD curves at various current densities from 1 to 20 A g−1 and the corresponding calculated rate performances are displayed in Fig. 6c. The PMoW–PDDA–RGO-based SC demonstrates a maximum specific capacitance of 177, 165.6, 151, 140 and 124 F g−1 at current densities of 1, 2, 5, 10 and 20 A g−1, respectively, which are superior to the previously reported results in the examples of Cs–PMo12/CNT (20 F g−1 at 2 A g−1),40 PMo12/CNTs (40 F g−1 at 1 A g−1),22 and PMo12/RGO (123 F g−1 at 1 A g−1).24 Cyclability is also a crucial factor for SCs. Fig. 6d exhibits an excellent cycling stability with retention of 94.6% of its initial capacitance after 1700 cycles at a high current density of 5 A g−1. The good electrochemical capability of the PMoW–PDDA–RGO composite predicts a possibility for practical SC application.
An EIS test (Fig. 7) was employed to further study the charge transfer process at electrode/electrolyte interfaces. The equivalent series resistance (from the x-intercept of the Nyquist plot) of the PMoW–PDDA–RGO composite is 1.67 Ω and that of RGO is 2.07 Ω. The smaller value for the former implies much better ion diffusion and a lower contact resistance for the PMoW–PDDA–RGO composite. This is benefited from the 3D porous interconnected network of the PMoW–PDDA–RGO composite.
In the two-electrode setup, the gravimetric specific energy density (E) and power density (P) of the symmetrical SCs are calculated based on the GCD curves from the following equations:
E = Csp × ΔV2/4 × (2 × 3.6) | (4) |
P = E × 3600/t | (5) |
The specific energy density of the PMoW–PDDA–RGO symmetric SC is 6.15 W h kg−1 with a power density of 250 W kg−1 and retains 4.31 W h kg−1 with a high power density of 5 kW kg−1. The symmetric supercapacitor device developed in this work exhibits improved performances compared to most of the PMo12-based symmetric supercapacitors previously reported, including PMo12/polypyrrole//PMo12/polypyrrole (1.44 W h kg−1 at 18.9 W kg−1),41 PMo12/CNTs//PMo12/CNTs (1.3 W h kg−1 at 50 W kg−1),22 and PMo12/RGO//PMo12/RGO (4.3 W h kg−1 at 250 W kg−1).24
Footnote |
† Electronic supplementary information (ESI) available: FT-IR spectrum, XRD, and 31P NMR analyses of PMo12, PW12 and PMoW. XPS of GO, RGO, and the PMoW–PDDA–RGO composite. N2 adsorption–desorption isotherms and pore size distribution of the PMoW–PDDA–RGO composite. CV curves of PMo12 and PW12 at 10 mV s−1. See DOI: 10.1039/c6ra15381j |
This journal is © The Royal Society of Chemistry 2016 |