Facile synthesis of ultra-small ruthenium oxide nanoparticles anchored on reduced graphene oxide nanosheets for high-performance supercapacitors

F. Z. Amir*a, V. H. Phamb and J. H. Dickersonb
aDepartment of Chemistry, Physics and Geology, Winthrop University, Rock Hill, SC 29733, USA. E-mail: amirf@winthrop.edu
bCenter of Functional Nanomaterials, Brookhaven National Laboratory, Upton, NY 11973, USA

Received 18th June 2015 , Accepted 3rd August 2015

First published on 3rd August 2015


Abstract

Herein, we report a facile, low cost, and environmentally friendly approach to prepare reduced graphene oxide–ruthenium oxide hybrid (RGO–RuO2) materials for supercapacitor electrode applications by in situ sol–gel deposition of RuO2 nanoparticles on the surface of graphene oxide (GO), followed by a reduction of GO in a strong alkaline medium at a low temperature. The combination of the sol–gel route and the reduction of graphene oxide at low temperatures resulted in ultrafine, hydrated amorphous RuO2 particles with sizes of only 1.0–2.0 nm, which uniformly decorated the surfaces of RGO sheets. The obtained RGO–RuO2 supercapacitor exhibited excellent electrochemical capacitive performance in a 1 M H2SO4 electrolyte with a specific capacitance more than 500 F g−1 at a current density of 1.0 A g−1 and high rate performance with the capacitance retention of 86% when the current density was increased 20 times, from 1.0 to 20.0 A g−1 in a two-electrode test cell configuration. The RGO–RuO2 system also showed good cycling stability with a capacitance retention of 87% after 2000 cycles. The excellent capacitive properties of RGO–RuO2 could be attributed to the uniform anchoring of ultra-small, hydrated amorphous RuO2 nanoparticles on the surface of RGO sheets, resulting in synergistic effects between them. The developed approach represents an exciting direction for enhancing the device performance of the graphene–metal oxide composite supercapacitors and can be used for designing the next generation of energy storage devices.


Introduction

Supercapacitors, a class of electrochemical energy storage devices with high power capacity, pulse power supply, exceptionally long cycling life, fast dynamics of charge propagation and low maintenance cost, are considered one of the most important devices for the next generation of energy storage.1–3 Bridging conventional capacitors and batteries, supercapacitors have several applications, such as energy back-up systems, portable devices, power tools, and hybrid electric vehicles.1–4 Supercapacitors can be classified into two categories based on the charge-storage mechanism: electrical double-layer capacitors (EDLCs) and pseudocapacitors. EDLCs achieve energy storage by forming a double layer of electrolyte ions on the surface of the conductive electrodes. Hence, high specific surface area and electrical conductivity of the electrode are crucial to ensure good performance of EDLCs.1,2 Carbon-based materials with high surface area, such as activated carbons, mesoporous carbons, carbon nanotubes, carbon aerogels, and recently graphene, are the most common materials in EDLCs. EDLCs composed of carbon electrodes have shown very high power density due to their fast kinetics on the adsorption or desorption of electrolyte ions; however, their relatively low energy density significantly limits their practical applications. In contrast to EDLCs, the capacitance of a pseudocapacitor comes from faradaic redox reactions at the electrode/electrolyte surface.4 Transition metal oxides and conducting polymers such as RuO2, MnO2, V2O5, polyaniline, and polypyrrole, are among the most common pseudocapacitive materials. Although pseudocapacitive materials could possess a specific capacitance that is 10–100 times higher than that of EDLCs, they usually suffer from relatively low power density and instability during cycling.5,6 The incorporation of pseudocapacitive materials into EDCL material matrices, creating hybrid structures, is the most common strategy to achieve both high power density and energy density as well as good cycling stability of supercapacitors.

Graphene, a one-atom-thick two dimensional (2D) single layer of sp2-bonded carbon, is considered as an ideal EDLC electrode material due to its extremely large surface area, extraordinarily high electrical conductivity, good chemical stability, and high mechanical strength.1–3 On the other hand, RuO2 has been the most extensively studied candidate for pseudocapacitor electrode materials, given its wide potential window, highly reversible redox reaction, remarkably high specific capacitance, high rate capacity and long cycle life.4,6 Combining the ideal EDLC electrode material with the most promising pseudocapacitive material is expected to create the best electrode material for supercapacitors. So far, several attempts have been made to combine these two kinds of materials.7–17 However, the capacitive performance of RGO–RuO2 nanocomposites has varied, ultimately depending on the dispersion morphology and crystallinity of RuO2 particles on RGO. To obtain high capacitive performance, RuO2 must be in the form of small sized, hydrated, amorphous nanoparticles, uniformly anchored on the surface of an RGO single sheet.6 Since the pseudocapacitance of RuO2 comes from the reversible faradaic redox reaction, the hydrated amorphous structure of RuO2 nanoparticles allows the redox reactions to occur not only at the surface (like crystalline RuO2 nanoparticles) but also within the inner parts of the RuO2 nanoparticles. The smaller the size of the RuO2 nanoparticles is, the higher the specific surface area is. Thus, more metal centers will be able to provide multiple redox reactions resulting in higher specific capacitance values.6,18

The uniform anchoring of RuO2 nanoparticles onto the surface of RGO sheets provides conduction paths that enable electrons to be easily transferred from the current collector to RuO2 nanoparticles and vice versa. So far, the synthesis of RGO–RuO2 can be classified into two strategies based on the RGO precursor. The first and most common approach is using RGO as a precursor in sol–gel or hydrothermal processes.7–10,12 In this approach, RGO is first prepared by the chemical reduction or thermal exfoliation/reduction of GO and the resultant material's use as a precursor in a sol–gel or a hydrothermal process. However, since RGO is usually in an aggregated or a restacked form, which is not exfoliated or well dispersed in common polar solvents (e.g. water, methanol, ethanol or isopropanol), the deposition of RuO2 nanoparticles on aggregated RGO sheets results in a loss of the effective surface area. More importantly, since RGO has fewer oxygen functional groups on the surface, which might act as nucleation and anchoring sites for the oxide formation, the density of RuO2 nanoparticles anchored on the surface of RGO is less. Therefore, RuO2 nanoparticles have the tendency to grow during the hydrothermal process, which is unfavorable for capacitive properties.15 Surfactants, like polyvinylpyrrolidone, polyelectrolytes, or poly (diallyldimethylammonium chloride), have been used as a stabilizer to improve the dispersion of RGO in water and the dispersion of RuO2 nanoparticles on RGO sheets.12,16 However, these stabilizers, which are absorbed on the surface of RGO–RuO2, are undesirable as a component of electrode materials and must be removed prior to the further use of RGO–RuO2. A second approach involves GO as a precursor in hydrothermal processes. In contrast to RGO, GO is completely exfoliated and dispersed in polar solvents and has abundant oxygen functional groups on the surface, providing numerous sites for RuO2 deposition. As a result, small and uniform RuO2 nanoparticles can decorate the surface of GO sheets. However, since GO is an insulating material, the hydrothermal process is usually carried out to reduce GO, recovering the electrical conductivity.14–16 Unfortunately, high temperature hydrothermal processes (180–200 °C) result in the recrystallization of RuO2, which is undesirable.15,16

In this study, we have designed a facile approach to prepare RGO–RuO2 hybrid composites through the in situ sol–gel deposition of RuO2 nanoparticles on the surface of GO, followed by a reduction of GO in a strong alkaline medium at low temperature. The combination of sol–gel techniques and the reduction of GO at low temperatures resulted in ultrafine, hydrated amorphous RuO2 nanoparticles of a diameter 1.0–2.0 nm that uniformly decorated the surfaces of RGO sheets. The electrochemical capacitive properties of RGO–RuO2 were characterized using a two-electrode test cell configuration in 1 M H2SO4 and 1 M Na2SO4 electrolyte solutions in the potential window of 1.0 and 1.5 V, respectively. The RGO–RuO2 exhibited excellent electrochemical capacitive performance in H2SO4 with a specific capacitance more than 500 F g−1 at a current density of 1.0 A g−1. This material also exhibited high rate performance with a capacitance retention of 86% when increasing the current density by 20 times, from 1.0 to 20.0 A g−1. RGO–RuO2 also displayed good cycling stability with a capacitance retention of 87% after 2000 cycles.

Experimental

Preparation of GO

GO was synthesized from expanded graphite by the modified Hummers method, described elsewhere.20 Briefly, 200 mL of concentrated H2SO4 was charged into a 1000 mL beaker equipped with a mechanical stirrer (Teflon impeller). Two grams of expanded graphite was gradually added under stirring to make a suspension. Then, 15 g of KMnO4 was slowly added. The temperature was then elevated to 35 °C, and the suspension was stirred for 2 h. The beaker was then chilled in an ice bath, and 500 mL of deionized water was slowly added to maintain a temperature below 70 °C. The mixture was stirred for 1 h and subsequently diluted with 3.0 L of deionized water. Twenty mL of H2O2 (30 wt%) was slowly added, and vigorous bubbles appeared as the color of the suspension changed from dark brown to yellow. The suspension was centrifuged and then washed with 1 M HCl solution four times to remove manganese compound residues. Subsequently, the H2SO4 was removed by copiously centrifuging and washing with deionized water. The obtained GO existed as a highly viscous solution.

Preparation of RGO–RuO2 hybrid

The as-prepared GO was diluted to a concentration of 0.5 mg mL−1 and sonicated for 10 min to obtain a homogenous solution. 622 mg of ruthenium chloride hydrate (RuCl3·xH2O, Ru content 40–49%) was then added into 200 mL of GO solution (0.5 mg mL−1) under mixing. Then, the suspension was neutralized with 1 M NaOH to a pH ∼7 and stirred for 12 h at room temperature to obtain GO–RuO2. Subsequently, the pH of the GO–RuO2 suspension was adjusted by adding 1 M NaOH to 12, and the suspension was aged at 90 °C for 12 h to deoxygenation of GO.21 The RGO–RuO2 suspension was repeatedly centrifuged and washed with deionized water and freeze-dried. The resulting RGO–RuO2 was denoted as the as-prepared RGO–RuO2. Finally, as-prepared RGO–RuO2 was annealed at 150 °C for 2 h to obtain annealed RGO–RuO2 (a-RGO–RuO2). To provide a materials standard for characterization, RGO was prepared following the above procedure without the additional ruthenium chloride.

Characterization

The morphologies of the a-RGO–RuO2 materials were characterized by scanning electron microscopy (SEM, JEOL 7600F) and transmission electron microscopy (TEM, JEOL 1400). The elemental mapping of a-RGO–RuO2 was provided by energy dispersive X-ray spectroscopy (EDS) and was performed on JEOL 7600F. Structural analysis was carried out on the Rigaku Ultima III X-ray diffractometer with CuKα radiation (λ = 0.15418 nm) at 40 kV and 44 mA. Thermogravimetric analysis (TGA) was performed under argon and air atmospheres at a heating rate of 5 °C min−1 (Perkin Elmer).

Electrode preparation and electrochemical measurements

The a-RGO–RuO2 electrodes were prepared by mixing a-RGO–RuO2 with carbon black and polytetrafluoroethylene (PTFE) in a mass ratio of 80[thin space (1/6-em)]:[thin space (1/6-em)]15[thin space (1/6-em)]:[thin space (1/6-em)]5 in ethanol to make a slurry. The mixture was sonicated for 5.0 min. The slurry was partially dried to make a paste, which was subsequently casted and pressed onto stainless steel and nickel foam (1.5 × 1.5 cm2) substrates that were used as current collectors. All casts were dried at 100 °C for 3 h. The mass of a-RGO–RuO2 was about 4.5–5.0 mg cm−2. Prior to the electrochemical measurement, the electrodes were soaked in each of the electrolytes for 12 h.

The capacitive performance of the a-RGO–RuO2 was characterized for a two-electrode configuration. Two identical electrodes, separated by a filter paper separator in 1 M Na2SO4 and 1 M H2SO4 electrolytes, were sandwiched in a supercapacitor test cell (EQ-STC split flat cell, MTI Corp.). Cyclic voltammetry (CV) measurements were performed on a potentiostat/galvanostat PARSTAT 2273 (Princeton Applied Research), and galvanostatic charge–discharge (GCD) tests were conducted on an Arbin battery tester BT2000 (Arbin Instrument) in the potential range of 0–1.5 V and 0–1.0 V for 1 M Na2SO4 and 1 M H2SO4, respectively. Electrochemical impedance spectroscopy (EIS) tests were performed over a frequency range from 0.01 Hz to 100 kHz at an open circuit potential with an AC perturbation of 10.0 mV. The mass specific capacitance, energy density, and power density were calculated according to the following equations:22,23

 
image file: c5ra11772k-t1.tif(1)
 
image file: c5ra11772k-t2.tif(2)
 
image file: c5ra11772k-t3.tif(3)
where Csp is the specific capacitance, I is the constant discharge current, Δt is the discharging time, m is the mass of two electrode, ΔV is the voltage drop upon discharging, E is the energy density, and P is the power density.

Results and discussion

The synthesis of a-RGO–RuO2 hybrids, as illustrated in Fig. 1, involved sol–gel techniques followed by the deoxygenation of GO. By adding the RuCl3·xH2O into the GO solution, the viscosity of the solution significantly increased because of the strong interactions between cationic Ru3+ and the oxygen functional groups (hydroxyl, epoxy and carboxyl groups) on the surface and the edge of the GO sheets. This led to abundant cross-linking between the GO sheets.24 The abundant oxygen functional groups of GO strongly interact with Ru3+, making them the active sites for nucleation and anchoring of the RuO2 nanoparticles during the sol–gel process. This results in uniform coating of RuO2 nanoparticles on the surface of GO sheets. Although the GO was slightly reduced during the sol–gel process,15 that reduction is insufficient to recover the electrical conductivity that is vital for effective capacitive electrode materials. The further reduction of GO was carried out in a strong alkaline medium at low temperature to prevent the growth and recrystallization of RuO2 nanoparticles. The reduction of GO in an alkaline medium is well-known as an effective and green approach to prepare RGO. More interestingly, unlike other reduction methods, the reduction of GO in an alkaline medium yields in highly dispersible RGO.21
image file: c5ra11772k-f1.tif
Fig. 1 Schematic illustration of the preparation procedure for the a-RGO–RuO2 hybrid.

After the alkaline reduction, the coagulated RGO–RuO2 particles were observed to be re-dispersible in water or ethanol by mild sonication. Annealing the as-prepared RGO–RuO2 at 150 °C for 2 hours optimally removed both absorbed and chemically bound water from the RuO2 nanoparticles to improve the capacitive performance of RGO–RuO2.6–10,25,26 Higher annealing temperatures may result in complete dehydration and recrystallization of RuO2 nanoparticles, leading to a capacitance decrease.

Fig. 2a–d shows SEM images of a-RGO–RuO2, revealing a three-dimensional porous structure consisting of interconnected flakes with pore sizes as large as ten microns. The RuO2 nanoparticles were densely and homogeneously anchored across the wrinkled RGO sheets. To confirm the uniform dispersion of the RuO2 nanoparticles on RGO sheets, EDS mapping was performed, as shown in Fig. 3b–d. Three elements, carbon (C), oxygen (O) and ruthenium (Ru), were observed in the map; the densities and the distribution of C, Ru and O matched the topology of the corresponding a-RGO–RuO2. This suggested a uniform distribution of RuO2 nanoparticles on RGO sheets. The elemental compositions of a-RGO–RuO2, as determined by EDS, were 46.18, 27.85 and 25.97 wt% for C, O and Ru, respectively.


image file: c5ra11772k-f2.tif
Fig. 2 (a–d) SEM images of a-RGO–RuO2 at different magnifications.

image file: c5ra11772k-f3.tif
Fig. 3 (a) SEM image of a-RGO–RuO2 and (b–d) SEM EDS mapping of (a).

The microstructure of a-RGO–RuO2 was further characterized by TEM. The low magnification of TEM micrographs in Fig. 4a showed the micron-sized wrinkled RGO sheets, which were densely coated by RuO2 nanoparticles. Although the suspension was sonicated for 1 minute during the TEM sample preparation, no unbound RuO2 nanoparticles were found, indicating strong interactions between RuO2 nanoparticles and RGO sheets. The high magnification TEM image in Fig. 4b revealed ultra-small RuO2 nanoparticles with the size of only 1.0–2.0 nm, disorderedly anchored on the surface of the RGO sheet, creating nanopores and nanochannels which may act as electrolyte reservoirs, facilitating the ions transportation during electrochemical redox reactions. The inset of Fig. 4c displays a TEM diffraction pattern of a-RGO–RuO2 consisting of diffuse ring diffraction patterns with no discrete reflections; this indicated that a-RGO–RuO2 has an amorphous structure.


image file: c5ra11772k-f4.tif
Fig. 4 (a & b) TEM images and the inset show TEM diffraction of a-RGO–RuO2.

Fig. 5 shows the X-ray diffraction (XRD) patterns of GO, RGO, as-prepared RGO–RuO2 and a-RGO–RuO2. The XRD pattern of GO showed a strong peak at 10.6°, corresponding to the (002) reflection with interlayer spacing of 8.32 Å. After the alkaline reduction, a broad peak was observed at 23°, corresponding to an inter-sheet spacing of 3.8 Å, which is distinctly smaller than the interspacing between GO sheets due to the removal of the oxygen functional groups on the surface of RGO.27 The XRD patterns of the as-prepared RGO–RuO2 and a-RGO–RuO2 do not exhibit peaks between 10 to 80°, implying that both the as-prepared RGO–RuO2 and the a-RGO–RuO2 are amorphous. Moreover, the disappearance of the broad RGO peak at 23° indicates that the anchored RuO2 nanoparticles on the surface of RGO sheets act as a spacer, effectively preventing the restacking of RGO sheets.


image file: c5ra11772k-f5.tif
Fig. 5 XRD patterns of GO, RGO, as-prepared RGO–RuO2 and a-RGO–RuO2.

The thermal stability of a-RGO–RuO2 in air and in an inert atmosphere was characterized by TGA as shown in Fig. 6. The TGA curve of a-RGO–RuO2 in argon atmosphere shows appropriately 8.7 and 14.6% mass losses at 150 and 600 °C, respectively. The first mass loss can be attributed to the evaporation of the absorbed water, whereas the second mass loss is ascribed to the removal of chemically bound water in hydrated RuO2 and the removal of labile oxygen functional groups of RGO, such as hydroxyl, epoxy and carbonyl groups.7,14,20,25 The significant mass loss of a-RGO–RuO2 at temperatures lower than 150 °C indicated that RuO2 exists in a hydrous form after thermal annealing, which is essential for enhancing the diffusion of cations inside the electrode material.6 The TGA of a-RGO–RuO2 in air showed almost an identical mass loss to that of a-RGO–RuO2 in an inert atmosphere at temperatures lower than 150 °C. However, the significant mass loss was observed in the temperature range of 150–350 °C, which can be mainly attributed to the oxidation of RGO under the catalytic influence of RuO2 nanoparticles.28 The residue of a-RGO–RuO2 was almost unchanged, approximately 72.6 wt%, at temperature higher than 400 °C, which can be assigned to the mass of RuO2. The RGO content in a-RGO–RuO2 could be estimated to be in a range of 12.9–18.8 wt%. The first value (12.9 wt%) was estimated from the difference of the residues of a-RGO–RuO2 at 600 °C in argon and air atmospheres whereas the second value (18.8 wt%) came from the difference in the residues of a-RGO–RuO2 at 150 °C and 600 °C in air atmosphere. Correspondingly, the mass ratio of RGO[thin space (1/6-em)]:[thin space (1/6-em)]RuO2 was about 1[thin space (1/6-em)]:[thin space (1/6-em)]3.8 to 1[thin space (1/6-em)]:[thin space (1/6-em)]5.6, which is closed to the theoretical input value of GO[thin space (1/6-em)]:[thin space (1/6-em)]RuO2 (1[thin space (1/6-em)]:[thin space (1/6-em)]4.0).


image file: c5ra11772k-f6.tif
Fig. 6 TGA of a-RGO–RuO2 in argon and air atmospheres.

The electrochemical capacitive behavior of a-RGO–RuO2 was evaluated by CV and GCD techniques. Fig. 7a & b shows CV curves of a-RGO–RuO2 at different scan rates in 1 M Na2SO4 (neutral) and 1 M H4SO4 (acidic) electrolytes, respectively. Taking advantage of the neutral electrolyte, the capacitive properties of a-RGO–RuO2 in Na2SO4 were characterized in the potential window of 0.0–1.5 V to improve the energy density.17,29 The CV curves of a-RGO–RuO2 in both electrolytes were nearly rectangular, and no redox peak was observed, indicating that the a-RGO–RuO2 had nearly ideal capacitive behavior. The absence of the redox peak can be explained by the very fast reversible redox reaction that occurred on the surface of RuO2 nanoparticles, which can be attributed to the ultra-small size of the hydrous RuO2 nanoparticles. According to Zhang et al.,6,19 the redox reaction can occur at a depth of 2 nm from the surface of the particle; considering the ultra-small size of the RuO2 nanoparticles, the redox reaction occurred on their surface. The current density values of the CV curves of a-RGO–RuO2 in H2SO4 were more than twice that of a-RGO–RuO2 in Na2SO4 at the same scan rate, indicating that the a-RGO–RuO2 has higher specific capacitance in H2SO4. The GDC curves of a-RGO–RuO2 in Fig. 7c & d were symmetrically triangular, which is another indication of ideal capacitive behavior. A small IR drop was observed at the beginning of the discharge curve, especially for a-RGO–RuO2 in H2SO4, implying the low internal resistance within the electrode.


image file: c5ra11772k-f7.tif
Fig. 7 (a and b) CV and (c and d) charge–discharge curves of a-RGO–RuO2 in Na2SO4 and H2SO4 electrolytes, respectively.

The specific capacitances of a-RGO–RuO2 calculated from the discharge curves were shown in Fig. 8a. The specific capacitances of a-RGO–RuO2 in H2SO4 were 509.4 and 439.2 F g−1 at current densities of 1.0 and 20.0 A g−1, respectively. These values are comparable to the best specific capacitances RGO–RuO2 ever reported, as shown in Table 1. Note that most of the reported, higher value, specific capacitances originate from measurements on three electrode test cell configurations, which do not accurately predict the more realistic capacitances measured from two electrode test cell configurations that mimic the physical configuration of a commercially packaged supercapacitor.23 In comparison to the a-RGO–RuO2 in H2SO4, the specific capacitances of a-RGO–RuO2 in Na2SO4 were much smaller, only 206.8 and 136.6 F g−1 at current densities of 1.0 and 20.0 A g−1, respectively. The lower specific capacitance of a-RGO–RuO2 in Na2SO4 is due to the shortage of protons within neutral Na2SO4, which are needed for the redox reaction of RuO2.6 The reduced capacitance of a-RGO–RuO2 in Na2SO4 primarily originated from the electrical double layer formed from the adsorption of ions at the surface of a-RGO–RuO2.


image file: c5ra11772k-f8.tif
Fig. 8 (a) Specific capacitance, (b) Nyquist plots, (c) cycling stability and (d) Ragone plots of RGO–RuO2 in Na2SO4 and H2SO4 electrolytes, respectively.
Table 1 Summary of specific capacitances of RGO–RuO2 collected from the literature
Material Synthesis method Structure RuO2 content Testing configuration Specific capacitance Ref.
RuO2-f-HEG Sol–gel + annealing at 350 °C, 2 h Crystalline 25 wt% Two-electrode 265 F g−1 at scan rate 10 mV s−1 8
RuO2/GNs Sol–gel + annealing at 150 °C, 6 h Amorphous 36 wt% Two-electrode 365 F g−1 at scan rate 5 mV s−1 10
RGO–RuO2 Sol–gel + annealing at 150 °C, 2 h Not reported 86.9 wt% Two-electrode 400 F g−1 at current density of 1.0 A g−1 9
GRA-6 Hydrothermal + annealing at 150 °C, 6 h Partial crystalline 45 wt% Three-electrode 471 F g−1 at current density of 0.5 A g−1 11
ROGSC Sol–gel + annealing at 150 °C, 2 h Partial crystalline 50.4 wt% Three-electrode 570 F g−1 at scan rate of 1 mV s−1 7
RuO2/GS Hydrothermal + annealing at 150 °C, 2 h Partial crystalline 52.7 wt% Three-electrode 551 F g−1 at current density of 1.0 A g−1 14
RuO2/RGO Hydrothermal Partial crystalline 75 wt% Three-electrode 497 F g−1 at current density of 0.5 A g−1 15
a-RGO–RuO2 Sol–gel + alkaline reduction of GO + annealing at 150 °C, 2 h Amorphous 72.6 wt% Two-electrode 509 F g−1 at current density of 1.0 A g−1 This work


Rate performance, one of the most important characteristics of a supercapacitor, was exceptional for the a-RGO–RuO2 system, particularly in H2SO4 (Fig. 8a). The respective capacitance retentions were 86.2% (H2SO4) and 66.1% (Na2SO4) when the current density increased from 1.0 to 20.0 A g−1. The uniform decorating of ultra-small RuO2 nanoparticles across the RGO sheets, insures fast charge carrier transport between the current collector and the surface of the RuO2. Moreover, a-RGO–RuO2's nanopores and nanochannels acted as electrolyte reservoirs, assuring a short diffusion length of ions within the electrode.

Electrochemical impedance spectroscopy was performed to extend our understanding of a-RGO–RuO2's electrochemical behavior. The Nyquist plots in Fig. 8b exhibited a typical arc in the high-frequency region and a straight line in the low-frequency region. The vertical shape of the straight lines in the low-frequency region indicated that a-RGO–RuO2 in both electrolytes closely resemble an ideal capacitor, consistent with the results from the CV and GDC. The arc and the Warburg-type line (the slope of 45° portion of the curve) of the a-RGO–RuO2 in H2SO4 was much shorter and smaller than those in Na2SO4, suggesting a lower charge transfer resistance and a more efficient electrolyte diffusion in H2SO4 than in Na2SO4.20 This result was reasonable because the diffusion of H+ protons via a hopping mechanism was much faster than the diffusion of Na+ cations. The equivalent series resistances (ESR), determined by extrapolating the vertical portion of the plot to the real axis, were only 0.62 and 1.55 Ω for a-RGO–RuO2 in H2SO4 and Na2SO4, respectively. The smaller ESR of a-RGO–RuO2 in H2SO4 explained its higher rate capacity.

The cycling stability, a critical factor for the practical implementation of supercapacitor electrodes, was evaluated using the galvanostatic charge–discharge technique at a current density of 2.0 A g−1. As seen in Fig. 8c, the cycling stability of a-RGO–RuO2 in Na2SO4 was slightly better than that in H2SO4. The capacitance of a-RGO–RuO2 noticeably decreased ∼5% in the first 200 cycles for both electrolytes. Then, the capacitance of a-RGO–RuO2 in H2SO4 electrolyte gradually decreased, whereas the capacitance in Na2SO4 was restored to 98% its initial capacitance after 350 cycles before slightly decreasing again. The capacitance retentions of a-RGO–RuO2 after 2000 charge–discharge cycles were 86.8% for H2SO4 and 94.5% for Na2SO4, indicating that a-RGO–RuO2 has good long-term electrochemical stability.

The Ragone plots of a-RGO–RuO2 are featured in Fig. 8d. Although the specific capacitance of a-RGO–RuO2 in Na2SO4 was less than half of that in H2SO4, the energy density a-RGO–RuO2 in Na2SO4 was comparable to that in H2SO4 due to extended working potential window, which was up to 1.5 V in the neutral electrolyte. The maximum energy densities of a-RGO–RuO2 of 16.7 and 15.0 W h kg−1 in H2SO4 and Na2SO4, respectively, were achieved at the current density of 1.0 A g−1, corresponding to the power density of about 1000 W kg−1. More interestingly, at the high power density of 10 kW kg−1, a-RGO–RuO2 delivered a reasonable energy density of 6.3 W h kg−1, indicating that a-RGO–RuO2 is a promising electrode material for high power applications.

Conclusions

We report a facile approach to prepare a-RGO–RuO2 for supercapacitor applications by an in situ sol–gel deposition of RuO2 nanoparticles on the surface of GO, followed by the reduction of GO in a strong alkaline medium at low temperature. The combination of the sol–gel route and the reduction of graphene oxide at low temperature resulted in ultrafine, hydrated amorphous RuO2 nanoparticles with the size of only 1.0–2.0 nm, which uniformly decorated the surfaces of the RGO sheets. The excellent specific capacitance (509 F g−1) obtained in H2SO4 electrolyte was the best value ever reported for RGO–RuO2, measured from a two-electrode test cell configuration.

The a-RGO–RuO2 exhibited a superior rate performance with capacitance retention of 86% by increasing the current density from 1.0 to 20.0 A g−1. The a-RGO–RuO2 also showed good electrochemical cycling stability of 86.8% capacitance retention after 2000 cycles. The excellent capacitive properties of a-RGO–RuO2 were attributed to the uniform anchoring of ultra-small, hydrated amorphous RuO2 nanoparticles on the surface of RGO sheets, resulting in synergistic effects between them. RGO sheets serve as conductive path for RuO2 nanoparticles while RuO2 nanoparticles decorating the surface of RGO sheets act as spacers to prevent the restacking of RGO sheets, creating nanopores and nanochannels and maximizing the surface area of a-RGO–RuO2. Considering the facile synthesis approach and the excellent capacitive properties, a-RGO–RuO2 has emerged as a promising electrode material for practical high-performance supercapacitor applications.

Acknowledgements

This work was supported in part by the U.S. Department of Energy, Office of Science, Office of Workforce Development for Teachers and Scientists (WDTS) under the Visiting Faculty Program (VFP). The research used resources of the Center for Functional Nanomaterials, which is a U.S. DOE Office of Science Facility, at Brookhaven National Laboratory under Contract No. DE-SC0012704.

References

  1. L. L. Zhang, R. Zhou and X. S. Zhao, J. Mater. Chem., 2010, 20, 5983 RSC.
  2. Y. Huang, J. Liang and Y. Chen, Small, 2012, 8, 1805 CrossRef CAS PubMed.
  3. H. J. Choi, S. M. Jung, J. M. Seo, D. W. Chang, L. Dai and J.-B. Baek, Nano Energy, 2012, 1, 534 CrossRef CAS PubMed.
  4. P. Simon and Y. Gogotsy, Materials for electrochemical capacitors, Nat. Mater., 2008, 7, 845 CrossRef CAS PubMed.
  5. B. Conway, V. Birss and J. Wojtowicj, J. Power Sources, 1997, 66, 1 CrossRef CAS.
  6. G. Wang, L. Zhang and J. Zhang, Chem. Soc. Rev., 2012, 41, 797 RSC.
  7. Z. S. Wu, D.-W. Wang, W. Ren, J. Zhao, G. Zhou, F. Li and H.-M. Cheng, Adv. Funct. Mater., 2010, 20, 3595 CrossRef CAS PubMed.
  8. A. K. Mishra and S. Ramaprabhu, J. Phys. Chem. C, 2011, 115, 14006 CAS.
  9. J. Zhang, J. Jiang, H. Li and X. S. Zhao, Energy Environ. Sci., 2011, 4, 4009 CAS.
  10. R. B. Rakhi, W. Chen, D. Cha and H. N. Alshareef, J. Mater. Chem., 2011, 21, 16197 RSC.
  11. Y. Chen, X. Zhang, D. Zhang and Y. Ma, J. Alloys Compd., 2012, 511, 251 CrossRef CAS PubMed.
  12. Y. Chen, X. Zhang, D. Zhang and Y. Ma, J. Alloys Compd., 2012, 541, 415 CrossRef CAS PubMed.
  13. N. Soin, S. S. Roy, S. K. Mitra, T. Thundat and J. A. McLaughlin, J. Mater. Chem., 2012, 22, 14944 RSC.
  14. N. Lin, J. Tian, Z. Shan, K. Chen and W. Liao, Electrochim. Acta, 2013, 99, 219 CrossRef CAS PubMed.
  15. J. Y. Kim, K.-H. Kim, S.-B. Yoon, H.-K. Kim, S.-H. Park and K.-B. Kim, Nanoscale, 2013, 5, 6804 RSC.
  16. J. Shen, T. Li, W. Huang, Y. Long, N. Li and M. Ye, Electrochim. Acta, 2013, 95, 155 CrossRef CAS PubMed.
  17. W. Wang, S. Guo, I. Lee, K. Ahmed, J. Zhong, Z. Favors, F. Zaera, M. Ozkan and C. S. Ozkan, Sci. Rep., 2013, 4, 4452 Search PubMed.
  18. L. Deng, J. Wang, G. Zhu, L. Kang, Z. Hao, Z. Lei, Z. Yang and Z.-H. Liu, J. Power Sources, 2014, 248, 407 CrossRef CAS PubMed.
  19. C. Zhang, H. Zhou, X. Yu, D. Shan, T. Ye, Z. Huang and Y. Kuang, RSC Adv., 2014, 4, 11197 RSC.
  20. V. H. Pham, T. Gebre and J. H. Dickerson, Nanoscale, 2015, 7, 5947 RSC.
  21. X. Fan, W. Peng, Y. Li, X. Li, S. Wang, G. Zhang and F. Zhang, Adv. Mater., 2008, 20, 4490 CrossRef CAS PubMed.
  22. L. L. Zhang, X. Zhao, M. D. Stoller, Y. Zhu, H. Ji, S. Murali, Y. Wu, S. Perales, B. Clevenger and R. S. Ruoff, Nano Lett., 2012, 12, 1806 CrossRef CAS PubMed.
  23. M. D. Stoller and R. S. Ruoff, Energy Environ. Sci., 2010, 3, 1294 CAS.
  24. S. Park, K.-S. Lee, G. Bozoklu, W. Cai, S. B. T. Nguyen and R. S. Ruoff, ACS Nano, 2008, 2, 572 CrossRef CAS PubMed.
  25. J. P. Zheng, P. J. Cygan and T. R. Jow, J. Electrochem. Soc., 1995, 142, 2699 CrossRef CAS PubMed.
  26. H. Kim and B. N. Popov, J. Power Sources, 2012, 104, 52 CrossRef.
  27. S. Park, J. An, I. Jung, R. D. Piner, S. J. An, X. Li, A. Velamakanni and R. S. Ruoff, Nano Lett., 2012, 9, 1593 CrossRef PubMed.
  28. K. Naoi, S. Ishimoto, N. Ogihara, Y. Nakagawa and S. Hatta, J. Electrochem. Soc., 2009, 156, A52 CrossRef CAS PubMed.
  29. H. Xia, Y. S. Meng, G. Yuan, C. Cui and L. Lu, Electrochem. Solid-State Lett., 2012, 15, A60 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.