Yi Chenga,
Jian Pana,
Martin Saundersb,
Shikui Yaoa,
Pei Kang Shenc,
Huanting Wangd and
San Ping Jiang*a
aFuels and Energy Technology Institute & Department of Chemical Engineering, Curtin University, Perth, WA 6102, Australia. E-mail: S.Jiang@curtin.edu.au
bCentre for Microscopy, Characterisation and Analysis (CMCA), The University of Western Australia, Clawley, WA 6009, Australia
cCollaborative Innovation Center of Sustainable Energy Materials, Guangxi University, Nanning 530004, China
dDepartment of Chemical Engineering, Monash University, Clayton, Victoria 3800, Australia
First published on 19th May 2016
The most significant challenge in the development of ultrafine oxide based supercapacitors is the poor microstructure stability due to the rapid agglomeration of the fine nanoparticles (NPs). Here, we developed novel amorphous MnOx structurally confined ultrafine NiO NPs (∼2.3 nm) supported on graphene, NiO@MnOx via a simple and facile self-assembly process with the assistance of microwave sintering. NiO@MnOx with a NiO:
MnOx weight ratio of 1
:
0.2 achieves a high capacitance of 966 F g−1 based on total electrode materials and 3222 F g−1 based on active materials at a discharge current density of 2 A g−1. Remarkably, the materials retain 100% capacitance after 2000 cycles at a charge and discharge current of 10 A g−1. In contrast, the durability of ultrafine NiO NPs without MnOx confinement is very poor, with 94% of the capacitance lost under identical cyclic conditions despite the initial high capacitance of 3696 F g−1. The substantially enhanced capacitance, durability and high rate capacity contribute to the formation of a nanoporous and amorphous MnOx layer on ultrafine NiO NPs, which provides the extraordinary structural confinement and enhances the mass transfer process. The results provide a new strategy to develop highly efficient and durable ultrafine nanosized electrode materials for supercapacitors.
Intensive work has been devoted to develop carbon–metal oxides hybrids with the aim to achieve high capacity for energy storage. Different nano-structured and hierarchy structures, such as nanoflowers,17,18 nanotubes,19 nanoplates16 and core–shell structures20 etc., have been developed to increase the efficiency. While conventional metal oxide structures normally exhibit mediocre specific capacitance and show low active material utilization due to the fact that ion insertion/extraction depth is limited in the range of few nanometers on the surface of electrode materials. Two dimensional transition metal oxides (2D TMOs) show promising opportunities due to the improvement in the utilization of active materials. For example, Zhu et al. prepared 2D nickel hydroxide nanosheets with micron-sized planar area and thickness of <2 nm and reported specific capacitance of 3650 F g−1 at 2 A g−1.21 Unfortunately, these 2D TMOs are hard to fabricate due to their diverse coordination styles.22
Ultrafine nano-sized (less than 3 nm) metal oxide NPs are promising electrode materials of supercapacitors because it is relatively easy to control the particle size. But there are few studies in the development of ultrafine nano-sized metal oxide NPs due to the fact that ultrafine NPs inevitably exhibits high surface free energy, high electrical resistance and sluggish charge and proton transfer,23 which could consequently promote the growth and aggregation of NPs and the degradation of the electrocatalytic activity of the materials. For example, Ni(OH)2 NPs with average size ∼4.8 nm can only retain ∼59% of the initial capacitance after 1000 cyclic voltammetry cycles at a scan rate of 20 mV s−1.14 Jahromi et al. found that NiO NPs with an average particle size of 8 nm only remain ∼60.6% of its initial capacitance after 1000 cyclic voltammetry cycles at 2 A g−1.24 Zhao et al.25 studied the capacitive performance of NiO supported graphene oxide with NiO NPs size of 5–7 nm and initial capacitance of 525 F g−1 and showed a high retention of 95.4% of initial capacitance after 1000 cycles but at a relatively low current density of 0.2 A g−1.
The growth and agglomeration of ultrafine NPs could be reduced by the nano- or structure confinement. It has been shown that Pt-based NPs can be stabilized via pore confinement or encapsulation.26–28 Shang et al. reported a facile and scalable wet-chemical process to prepare graphene-nanosheet-supported Pt NPs covered by mesoporous silica (mSiO2) layers. The chemically and thermally stable mSiO2 layer not only prevented the aggregation and restacking of graphene nanosheets resulted from the π–π stacking interactions, but also provided spaces for confining metal NPs.29 Galeano et al. synthesized Pt NPs with size of 3–4 nm encapsulated within the hollow graphitic spheres (HGS) with a specific surface area exceeding 1000 m2 g−1 and showed good long-term stability and high activity.26 The increased durability of Pt NPs was attributed to the structural confinement of HGS to suppress the agglomeration.
Here, we developed a new strategy to structurally confine ultrafine NiO NPs of 2.3 nm with an amorphous MnOx thin layer homogenously supported on graphene sheets, NiO@MnOx. The best results were obtained on NiO@MnOx NPs with NiO to MnOx weight ratio of 1:
0.2, achieving a high specific capacitance of 3222 F g−1 (based on the active materials), high material utilization, outstanding rate capability and columbic efficiency as well as excellent durability.
PEI functionalized graphene (30 mg) was first ultrasonicated in 100 mL EG solution for 1 h, followed by the addition of 33.8 mg nickel(II) acetylacetonate. The dispersion was ultrasonicated for 15 min and then stirred for another hour before being placed in a microwave oven (1000 W) and heated for 4 min, followed by stirring overnight. The solution was then filtered using a nylon filter membrane and washed for several times. The designed loading of NiO on graphene was 25 wt%, and the as-prepared catalysts were denoted as NiO.
NiO (10 mg) was dispersed in 40 mL Milli-Q water under ultrasonic for 30 min and appropriate amount of KMnO4 solution (1 mg mL−1) was added. The dispersion was stirred at room temperature for 30 min followed by adding 40 mL ethanol. The mixture then were refluxed at 100 °C in an oil bath for 2 h and filtered using a nylon filter membrane (0.2 μm) and washed for several times with ethanol. The loading of MnO2 formed was controlled as 5, 10 and 25 wt% on graphene. This resulted in the synthesized NiO@MnOx with NiO to MnOx weight ratio of 1:
0.2, 1
:
0.4 and 1
:
1 supported on graphene, which was denoted as NiO@MnOx (1
:
0.2), NiO@MnOx (1
:
0.4) and NiO@MnOx (1
:
1), respectively. For comparison, 25 wt% MnOx on PEI-functionalized graphene was prepared in the same way and the product was denoted as MnOx.
NiO, MnOx or NiO@MnOx on graphene (10 mg) was ultrasonically mixed in 2 mL of Nafion solution to form a homogeneous ink, followed by pasting 100 μL of catalyst ink onto the surface of a carbon electrode with diameter of 1 cm−2. The total electrode material loading was 0.5 mg cm−2. Pt foil (3.0 cm2) and saturated camel electrode (SCE) electrodes were used as the counter and reference electrodes, respectively. All potentials in the present study were given versus SCE reference electrode and all the experiment were performed in a 2 M KOH aqueous solution. Cyclic voltammetry (CV) and galvanostatic charge–discharge cyclic tests were conducted with different scan rates and current densities, using a Princeton potentiostat (Versastat 3, USA) in a standard electrochemical cell. Electrochemical impedance spectroscopy (EIS) measurements were conducted in the frequency range from 0.01 Hz to 100 kHz at open circuit potential with signal amplitude of 5.0 mV. The stability of electrode materials was tested at a charge–discharge of 10 A g−1 for 2000 cycles, and the electrodes after cycling test were soaked in DI water and then collected for STEM and elemental mapping analysis. The specific capacitance Cs was calculated from the CV curves and galvanostatic charge discharge curves.
3C2H5OH + 4KMnO4 → 3HC2H3O2 + 4MnO2 + 4KOH + H2O | (1) |
![]() | ||
Fig. 1 Synthesis of NiO@MnOx nanostructures supported on PEI-functionalized graphene through self-assembly route assisted by microwave sintering. |
After the deposition of MnOx, the zeta potential changed from 26.4 mV for NiO to 8.4, −2.4 and −17.5 mV for the deposition of 5%, 10% and 25% MnOx, respectively, indicating the self-assembly of negatively charged MnO4− ions.
![]() | ||
Fig. 2 (A) The XRD patterns and (B) high resolution XPS patterns of Ni 2p and Mn 2p regions of NiO and NiO@MnOx supported on graphene. |
The nature of the NiO and NiO@MnOx nanostructures was further investigated by XPS (Fig. 2B). The Ni 2p XPS spectrum possesses two shakeup satellites at 855.9 eV (Ni 2p3/2) and 873.4 eV (Ni 2p1/2) for NiO@MnOx (1:
0.2), about 0.1 eV lower than those for NiO with BE of 856.0 eV and 873.5 eV respectively. As the loading of NiO
:
MnOx reaches 1
:
1, the Ni 2p3/2 and Ni 2p1/2 peaks are observed at 855.6 eV and 873.1 eV, which is about 0.4 eV lower than that of NiO. The Mn 2p XPS spectrum displays two characteristic peaks at 642.9 eV and 654.6 eV, corresponding to Mn 2p3/2 and Mn 2p1/2 spin–orbit peaks for NiO@MnOx (1
:
0.2), about 0.7 eV higher than those of MnOx. The BE of Mn 2p3/2 and Mn 2p1/2 is 642.6 eV and 654.3 eV for NiO@MnOx (1
:
1), about 0.4 eV higher than those of MnOx. The negatively BE shift of Ni 2p and the positive BE shift of Mn 2p of NiO@MnOx indicate the electron donation from Mn to Ni.
Fig. 3 shows the TEM images of the NiO and NiO@MnOx. The graphene sheet was homogenously covered by high density fine NPs (Fig. 3A) as compared with the pristine graphene (Fig. S2†). No obvious large particles observed, indicating that the self-assembly method is an effective method to obtain ultrafine nanostructures. The high resolution TEM image discloses that the average size of NiO NPs is 2.3 nm (Fig. 3A). This is larger than the crystallite size of 1.3 nm obtained from the XRD. The selected-area electron diffraction (SAED) pattern of the NiO reveals well-defined diffraction rings (Fig. 3A), suggesting the polycrystalline nature of NiO. This is consistent with the HRTEM image, which shows the NiO with a lattice space 0.21 nm (the inset, Fig. 3A).
![]() | ||
Fig. 3 TEM micrographs and histograms of (A) NiO, (B) NiO@MnOx (1![]() ![]() ![]() ![]() ![]() ![]() |
In the case of NiO@MnOx (1:
0.2), average particle size of NPs is 6.1 nm and uniformly distributed on graphene (Fig. 3B). As shown by the XRD results (Fig. 2A), the size of NiO NPs does not change after the deposition of MnOx layer, hence the large particle size of NiO@MnOx (1
:
0.2) indicates that the NiO NPs are covered by a thin layer of amorphous MnOx. Based on the average size of 2.3 nm NiO NPs, thickness of the amorphous MnOx layer deposited can be estimated to be ∼1.9 nm. The inset HRTEM in Fig. 3B further indicates that the MnOx appears to be porous, exhibiting large amount of nanopores of ∼0.5 nm in diameter. With the increase of the MnOx loading, the size of the NiO@MnOx increases (Fig. 3C and D), indicating the formation of NiO@MnOx nanostructures with thick MnOx layer. In the case of NiO@MnOx (1
:
0.4), the size of NPs is 9–13 nm, and the size of NiO@MnOx (1
:
1) nanostructure is larger than 20 nm and is interconnected (Fig. 3D). The continuous and interconnected NiO@MnOx nanostructure would make it hard to calculate the thickness of the MnOx films. Nevertheless, based on the fact that the size of NiO NPs does not change during the deposition of MnOx, the thickness of MnOx could be estimated to be 3–4 nm for NiO@MnOx (1
:
0.4) and 8–10 nm for NiO@MnOx (1
:
1), respectively. The SAED diffraction patterns of NiO@MnOx are identical with fcc NiO and no diffraction rings attributed to MnOx were found (Fig. 3B–D). Both XRD and HRTEM results indicate that deposited MnOx layer is amorphous and nanoporous. A schematic structure model of NiO@MnOx NPs is shown in Fig. 1.
Fig. 4 shows the elemental mapping using EDS coupled with STEM of NiO@MnOx with different NiO/MnOx ratios. The STEM images of the NiO@MnOx (1:
0.2) shows the ultrafine NiO NPs covered by a thin MnOx shell, forming porous nanostructures (Fig. 4A–C). The EDS mapping displays all three elements, Mn, Ni and O, which further indicates that the NiO NPs are covered by a thin layer of amorphous MnOx. With the increase of the MnOx loading, the EDS peak intensity significantly increases and at the same time, the intensity of the color associated with Mn also increases, indicating an increase of MnOx thickness on the NiO NP surface (Fig. 4D). In the case of NiO@MnOx (1
:
0.4), the number of interconnected nanostructures increase (Fig. 4B). As the NiO/MnOx ratio reaches to 1
:
1, the structure shows NiO NPs embedded in a thick and continuous amorphous MnOx layer (Fig. 4C).
The nitrogen sorption isotherm profiles of NiO and NiO@MnOx show a shape similar with type-IV with a H3 hysteresis loop associated with slit-shaped pores (Fig. 4E).34,35 The knee-point at low P/P0 of isotherm profiles is due to the formation of micropores likely resulted from the inherent tunnels or pores of amorphous MnOx. The BET surface area increase from 149.8 m2 g−1 for NiO to 280.1, 340.8 and 430.9 for NiO@MnOx (1:
0.2), NiO@MnOx (1
:
0.4) and NiO@MnOx (1
:
1), respectively. The increase of the BET surface area could be due to the connection and formation of hierarchy microstructures (the increase of hysteresis loop) and the micropores inherent from the porous structure of manganese oxide. Fig. 4F shows the formation of NiO@MnOx nanostructures with different NiO/MnOx ratios on graphene.
With the increase of NiO/MnOx ratio, the intensity of anodic and the cathodic peaks decreased because the covering of the NiO with a thicker layer of MnOx would increase the ionic insertion length and leads to a lower active material utilization. The anodic and cathodic peaks were symmetric, suggesting excellent reversibility for both NiO and NiO@MnOx electrodes. The minimal changes in the shape of the CV curves as the scan rate increases from 10 to 100 mV s−1 reveal the excellent electron conductivity of NiO and NiO@MnOx supported on graphene (Fig. S3†). The anodic and cathodic peaks, however, are positively shifted with increasing scan rate due to the internal resistance of the electrode. The anodic peak of NiO shifted by 70 mV from 0.28 V at the scan rate of 10 mV s−1 to 0.35 V at 100 mV s−1, which is higher than a positively shift of 60, 55 and 30 mV for NiO@MnOx (1:
0.2), NiO@MnOx (1
:
0.4) and NiO@MnOx (1
:
1) measured under the same conditions. This indicates the coverage of MnOx on the NiO nanostructures surface improve the reversibility and rate capacity.
The specific capacitance, Cs of NiO calculated from the CV curves at scan rate of 10 mV s−1 is 865 F g−1 based on the total mass of the electrode material (Ctotal, NiO + graphene) and 3460 F g−1 based on the active material NiO (Cactive). The Ctotal and Cactive is 1037 F g−1 and 3456 F g−1, respectively, for NiO@MnOx (1:
0.2). However, with the increased loading of MnOx, the Ctotal and Cactive decreased to 807 and 2305 F g−1 for NiO@MnOx (1
:
0.4), and 297 and 594 F g−1 for NiO@MnOx (1
:
1), respectively. The similar Cactive between NiO and NiO@MnOx (1
:
0.2) indicate that the deposition of a thin amorphous MnOx shell (∼1.9 nm) would not impede the fast electron transport and ion-diffusion to reach the active NiO NPs. With the increase in the scan rate, the Ctotal decreases gradually (Fig. 5B), which can be attributed to the inaccessible of the electrolytic ions (i.e., OH−) to the active NiO NPs for the charge storage at high scan rates. In the case of NiO NPs, Ctotal decreases to 252.7 F g−1 at scan rate of 200 mV s−1, which is only 29% of 865 F g−1 measured at scan rate of 10 mV s−1. The Ctotal of NiO@MnOx (1
:
0.2), NiO@MnOx (1
:
0.4) and NiO@MnOx (1
:
1) are 742.8, 588.7, and 230.8 F g−1, respectively, at scan rate of 200 mV s−1, which is 71.6%, 72.7% and 77.5% of that at scan rate of 10 mV s−1, much higher than 29% observed on NiO NPs. The much higher reduction in supercapacitance at high scanning rates for the reaction on NiO indicates that the rate capability of the NiO@MnOx electrodes is significantly higher than that of NiO electrode.
The electrochemical properties of NiO and NiO@MnOx were further characterized by galvanostatic charging and discharging method (Fig. 5C). The Cs of the electrodes was calculated from the discharge curve. The Ctotal and Cactive of the electrode material is 924 and 3696 F g−1 for NiO, 966 and 3222 F g−1 for NiO@MnOx (1:
0.2), 711 and 2031 F g−1 for NiO@MnOx (1
:
0.4), 263 and 526 F g−1 for NiO@MnOx (1
:
1), respectively, measured at a current density of 2 A g−1. For the reaction on NiO without deposition of amorphous MnOx, the Cs decreases rapidly with the increase of current density (Fig. 5D and S4†). Ctotal obtained at 40 A g−1 is 136 F g−1, which is 14.8% of 924 F g−1 measured at a current density of 2 A g−1. This indicates that pristine NiO NPs without the amorphous MnOx layer have a very low rate capability, which is consistent with that reported in the literature. For example, Liu et al. reported that the Cs of Ni(OH)2 NPs with average size ∼4.8 nm supported on reduced graphene oxide sheets decreased from 1578 F g−1 at 1 A g−1 to 905 at 10 A g−1, a reduction of 43% in capacitance.14
On the other hand, in the case of NiO@MnOx electrode materials, Ctotal measured at a current of 40 A g−1 is 713, 585 and 214 F g−1 for NiO@MnOx (1:
0.2), NiO@MnOx (1
:
0.4) and NiO@MnOx (1
:
1), respectively, which is 73.8%, 82.3% and 81.3% of that at current density of 2 A g−1 (Fig. 5D). This shows that the NiO@MnOx structure can retain 70–80% of its initial value when the discharge current density increases by 20 times, significantly better than 14.8% measured on NiO. The results show that the NiO@MnOx structure with a thin layer of amorphous MnOx can efficiently utilize the active materials even at high current density (high power demand). More importantly, the columbic efficiency (Cs discharge/Cs charge × 100%) of the NiO@MnOx is significantly higher than that of NiO. For example, the columbic efficiency is 100% for NiO@MnOx (1
:
0.2) at 40 A g−1, which is higher than 87.5% of NiO measured under identical conditions.
The impedance responses of NiO are characterized by a semicircle at high frequencies and a low frequency linear tail attributed to the ion-diffusion process between Warburg diffusion and ideal capacitive ion diffusion (Fig. 5E).11,16,39 The charge transfer resistance (Rct) associated with the high frequency arc of NiO is significantly higher than that of NiO@MnOx, revealing the decrease of charge transfer resistance after covering the NiO with amorphous MnOx. Most interesting, the NiO@MnOx (1:
0.2) exhibits the low frequency tail with a slope of ∼60°, as compared with a slope of ∼45° for the reaction on NiO@MnOx (1
:
0.4), NiO@MnOx (1
:
1) and NiO. This indicates the lower Warburg impedance and better ion diffusion of NiO@MnOx (1
:
0.2),11,16 consistent with its high performance.
The capacitance behavior of MnOx supported on graphene was also studied and the results are shown in Fig. 5F. The CV curve shows a rectangular shape, typical characteristics of double layer capacitance. Based on the discharge curve, the Ctotal and Cactive is 12.4 and 49.6 F g−1, lower than 110 F g−1 at 2 A g−1 reported on MnO2 nanosheets dispersed on single walled carbon nanotubes.40 The Cs of MnOx is substantially lower than that of NiO@MnOx, indicating that the high Cs of NiO@MnOx is mainly due to the NiO NPs and the contribution of deposited amorphous MnOx layer to the overall capacitance performance of NiO@MnOx is very low and can be neglected.
The microstructure of NiO, NiO@MnOx (1:
0.2) and NiO@MnOx (1
:
0.4) after cyclic test was analyzed by HAADF-STEM, and the results are shown in Fig. 7. In the case of NiO, a significant change of morphology was observed. NiO NPs aggregated into large particles with size around 12–50 nm (average ∼27.4 nm) (Fig. 7A), increased by an order of magnitude as compared to the initial particle size of 2.3 nm. In the case of NiO@MnOx, the change in the microstructure is much smaller. After the 2000 charge/discharge cycles, NiO NPs in the NiO@MnOx matrix can be clearly identified as indicated by the significantly increase intensity of the color associated with Ni (green in this case), different to that before the test (Fig. 4A). This indicates that originally embedded NiO NPs may extrude through the amorphous MnOx. In the case of NiO@MnOx (1
:
0.2), NiO NPs also grew to 2.0–6.5 nm (average ∼ 3.6 nm, Fig. 7B). However, NiO NPs are still clearly separated and the agglomeration of embedded NiO NPs is substantially smaller than that observed in the case of NiO. This evidently demonstrates the effectiveness of the confinement effect of the amorphous MnOx on the growth and agglomeration of NiO. Similar microstructure evolution was also observed for NiO@MnOx (1
:
0.4) (Fig. 7C). In this case, the embedded NiO NPs size is in the range of 2.5–9 nm (average ∼ 4.4 nm).
![]() | ||
Fig. 8 Comparison of specific capacitance based on active material of NiO@MnOx (1![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() ![]() |
The high performance of NiO@MnOx (1:
0.2) is clearly due to the ultrafine NiO NPs and this is supported by the initially high capacitance of 3696 F g−1 of ultrafine NiO NPs with size ∼2.3 nm supported on graphene. The ultrafine NiO NPs would allow short and efficient ion pathways for the charge and discharge process, leading to a high utilization of active material and favorable reaction kinetics. However, without the confinement of amorphous MnOx, NiO NPs grow and agglomerate quickly during the charge/discharge cycles, forming large particles with size of 12–50 nm (Fig. 7A). This explains the rapid loss of the capacitance of NiO NPs during the durability test (Fig. 6). The deposition of a thin layer of amorphous MnOx, such growth and agglomeration of embedded NiO NPs is significantly alleviated by the confinement of amorphous MnOx, as indicated by the 100% retention of the capacitance after the durability test and substantially grain growth and aggregation of active NiO NPs of NiO@MnOx (Fig. 6 and 7). Fig. 9 shows the structural model of agglomeration NiO NPs and structural confinement effect on NiO@MnOx nanostructure.
![]() | ||
Fig. 9 (A) Aggregation of ultrafine NiO NPs, and (B) structure confinement of the ultrafine NiO NPs core with amorphous MnOx shell. The inset picture is the structure model of NiO@MnOx. |
The electricity storage and discharge of NiO based electrode materials is through the following reactions in alkaline solution:
Ni(OH)2 + OH− ↔ NiOOH + H2O + e | (2) |
Thus, in the case of ultrafine NiO NPs, the rate capacity would strongly depend on the OH− concentration at the surface. With the increase of scanning rate or charge/discharge currents, capacitance of NiO decreased significantly (Fig. 5B and D), due to the significantly depletion of OH− concentration in the surface region. In the case of NiO@MnOx electrode materials, there is significantly enhanced rate capacity and reversibility, indicated by the negatively shifted ΔE for the redox reaction on NiO@MnOx and the much smaller reduction in Cs with the increase of discharge current density, as compared with NiO NPs.
Manganese oxides have been widely studied for energy storage and molecular sieve due to their inherent nanopores and interconnected channels, based on the MnO6 octahedral building blocks.46 The much higher rate capacity and reversibility of NiO@MnOx electrode materials clearly indicate that the presence of a nanoporous and amorphous MnOx layer substantially reduces the depletion of the reactant concentration, OH− at the surface of embedded NiO NPs, as compared to the free standing NiO NPs, an indication of enhanced mass transfer through the nanoporous structure of the deposited amorphous MnOx. Tagliazucchi and Szleifer proposed that ions transportation in nanopores would be enhanced driven by electrical field.47 Liu et al. revealed that the shorter the nanopore and/or the smaller its radius, the faster the osmotic flow.48 This implies that mass transfer rate within the nanochannels would be enhanced,49,50 and the nanoporous and amorphous MnOx layer could act as a “ion pump” to enhanced the mass transfer rate.51 This may explain the much less reduction in the capacitance of NiO@MnOx electrode materials as a function of the scan rate and charge/discharge currents, as compared to that of NiO. The negatively shift of Ni 2p peaks and the positively shift of Mn 2p peaks for NiO@MnOx (see Fig. 2B) also indicate that the MnOx could donation electron to the NiO and facilitate the charge and discharge of embedded NiO NPs.52 The results of this study demonstrate that the presence of microporous MnOx not only substantially inhibits the agglomeration of NiO NPs, but also enhances the mass transfer of the reactants for the charge/discharge, leading to highly active, durable, high utilization and high rate capacity of ultrafine NiO active materials for supercapacitors.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra04880c |
This journal is © The Royal Society of Chemistry 2016 |