Hierarchical porous nitrogen doped reduced graphene oxide prepared by surface decoration–thermal treatment method as high-activity oxygen reduction reaction catalyst and high-performance supercapacitor electrodes

Zhijun Jiaa, Baoguo Wangb, Yi Wang*a, Tao Qia, Yahui Liua and Qian Wangc
aNational Engineering Laboratory for Hydrometallurgical Cleaner Production Technology, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China. E-mail: wangyi@ipe.ac.cn; Fax: +86 10 82544848 802; Tel: +86 10 82544967
bDepartment of Chemical Engineering, Tsinghua University, Beijing 100084, China
cBeijing Key Laboratory of Ionic Liquids Clean Process, Key Laboratory of Green Process and Engineering, State Key Laboratory of Multiphase Complex Systems, Institute of Process Engineering, Chinese Academy of Sciences, Beijing 100190, China

Received 12th April 2016 , Accepted 8th May 2016

First published on 10th May 2016


Abstract

Nitrogen doped (N-doped) porous reduced graphene oxide (rGO) is successfully obtained by a two-step method, which includes a surface finishing of graphene oxide (GO) followed by thermal treatment. Distinct from many other methods, such as heat treatment, chemical vapor deposition or plasma treatment, a special atmosphere (NH3) and special equipment is avoided using this method. The nitrogen atoms are introduced into the reduced graphene oxide by this method according to the X-ray photoelectron spectroscopy (XPS) results. Based on the high specific surface area, high porosity, and doped heteroatoms, the as-prepared N-doped reduced graphene oxide shows excellent electrocatalytic activity towards the oxygen reduction reaction (ORR) and high capacitance, compared with a pristine reduced graphene oxide sample. Therefore, this work demonstrates a good example of both electrocatalytic activity and capacitance properties of N-doped reduced graphene oxide prepared by the surface decoration of GO followed by heat treatment method, which opens the door for creating functional graphene materials with highly promising applications in high-performance renewable energy conversion and storage devices.


1. Introduction

Graphene, a monolayer of sp2-bonded carbon atoms densely packed into a two-dimensional honeycomb crystal lattice, is recognized as an exciting material for its various superior properties such as high electrical and thermal conductivities, good transparency, great mechanical strength, inherent flexibility and huge specific surface area.1–4 Since the first report in 2004, tremendous work has been carried out to apply graphene into various applications by adjusting its physicochemical properties. Chemical functionalization of graphene oxide, GO, is an effective approach to tailor the properties of graphene, since enriched reactive oxygen functional groups in GO can provide ample covalent bonding sites for the chemical functionalization.5–7 Moreover, the most widely introduced heteroatom is nitrogen and it is known as N-doped graphene, which behaves differently from pristine graphene, and has been explored for many applications.8–12 In supercapacitor applications, it is reported that N-doped graphene has an improved specific capacitance compared to the pristine material.8–15 Furthermore, it also has considerable electrocatalytic activity for the ORR.7

Methods recently reported to synthesize N-doped graphene, include heat treatment under ammonia gas atmosphere at high temperature, chemical vapor deposition, plasma treatment and hydrothermal method.7–18 But all these methods have their inherent shortcomings. For example, no matter heat treatment, chemical vapor deposition or plasma treatment, special atmosphere (NH3) or special equipment is needed.8,10 For hydrothermal method, a long preparation time is required, usually over 48 hours, and the aggregation of reduced graphene oxide is hard to avoid and dramatically decreases the surface area, leading to a poor specific capacitance.7,17 All these are obstacles for high quality and mass productions of N-doped graphene. Therefore, a new strategy is needed to produce high quality N-doped graphene with a large scale.

In this work, we detailed a two-step method for doping nitrogen into graphene sheets with the use of melamine as the nitrogen source. At the meantime, the electrocatalytic activity to ORR and the capacitance performance of N-doped rGO were tested. It manifested that this novel strategy could produce high quality N-doped rGO and is suitable for mass production.

2. Experimental

2.1. Preparation of pristine rGO and N-doped rGO

Graphite oxide (GO) was synthesized using a modified Hummers method. 2 g of graphite and 1 g of sodium nitrate (NaNO3) were mixed in 50 mL sulfuric acid (H2SO4) by rigorous stirring for 3 hours in an ice bath at 0 °C. Then 8 g of potassium permanganate (KMnO4) was added slowly into the suspension, and the suspension was stirred until its color turned to dark gray. Next, 100 mL of DI water was slowly added to the suspension, and the suspension was maintained at 98 °C for 1 h. To dilute the suspension, 300 mL of DI water was added and followed by a 20 mL H2O2 addition to remove the residual KMnO4. Finally, a bright yellow graphite oxide solution was obtained. The as-prepared GO was washed with DI water several times, followed by a drying step to obtain graphite oxide power. The next step was to obtain GO solution from the graphite oxide powder by sonication. Typically, 2.5 g of the graphite oxide powder was added to 200 mL of DI water, and this solution was sonicated at 300 W for 3 h. The sonicated solution was centrifuged for 30 min at 2000 rpm to remove the remaining graphite oxide powder. The supernatant solution was about 10 mg mL−1 GO.

The pristine rGO, designated as rGOA, was prepared by reducing 200 mL of the 6 mg mL−1 GO solution with 6 g of ascorbic acid at 95 °C for 3 hours, as route (A) in Scheme 1. When the reduction process was completed, black rGOA was filtered and washed with DI water and ethanol several times. Finally, rGOA powder was obtained by vacuum drying at 70 °C overnight.


image file: c6ra09402c-s1.tif
Scheme 1 Schematic illustrations showing the synthetic routes of rGOA and MrGOth.

The N-doped rGO, designated as MrGOth, was prepared by first functionalizing GO sheets with melamine resin (M-GO), followed by a heat treatment under a vacuum, as route (B) in Scheme 1. For surface modification with melamine resin, melamine resin solution was firstly prepared by dissolving 2.5 g of melamine and 5 mL of 37 wt% formaldehyde solution into 60 mL of DI water at 70 °C with constant magnetic stirring in a flask. After the solution turned transparent, 20 mL of the prepared GO solution was added to melamine resin solution and stirred at 98 °C for 3 hours under continuous stirring. During the stirring, the condensation reaction between melamine resin and GO sheets took place, and bright silvery M-GO powder precipitated. After the reaction completed, M-GO was filtered and thoroughly washed with DI water and ethanol several times. Finally, M-GO powder was obtained by vacuum drying at 70 °C overnight. MrGOth was obtained by a multistep heat treatment of M-GO (120 °C for 20 min, 800 °C for 2 h and 400 °C for 2 h) with a ramping rate of 5 °C min−1.

2.2. Material characterization

The structures of samples were examined by X-ray diffraction (XRD, X'Pert-PRO MPD) using Cu Kα radiation with a wave length of 1.54 Å. The morphology was examined using scanning electron microscopy (SEM, JEOL JSM-6700F) and transmission electron microscopy (TEM, JEOL JEM-2100F), respectively. The specific surface area was measured using the Brumauer–Emmett–Teller (BET) method based on the nitrogen adsorption–desorption isotherm at 77 K on a Micrometritics ASAP2020 sorption analyzer. Surface chemical information was collected by X-ray photoelectron spectroscopy (Thermo ESCALB 250Xi).

2.3. Electrode preparation

To prepare the working electrodes, 2 mg of MrGOth was dispersed in 1 mL of 5 wt% Nafion aqueous solution. Then the mixture was ultrasonicated for 30 min to obtain a homogeneous catalyst ink. 0.5 μL of 2 mg mL−1 MrGOth dispersion was dropped onto a mirror like polished glassy carbon electrode with a diameter of 5 mm and dried in air at 70 °C. The rGOA and Pt/C (20%) was deposited on the electrode under the same procedure.

2.4. Electrochemical characterization

All the electrochemical measurements were conducted on an electrochemical workstation system (CHI760D, Chenhua, Shanghai) with a three-electrode cell. Pt foil and Hg/HgO (1 mol L−1 KOH) electrodes were used as the counter and reference electrodes, respectively. An aqueous solution of KOH (0.1 mol L−1) was used as electrolyte for rotating disk electrode (RDE) voltammogram measurements. Before every test, an O2 flow was used through the electrolyte for at least 30 min. For the rotating disk electrode test, the same amount catalysts were loaded on a rotating disk electrode. The polarization curves for ORR were conducted with a scan rate of 10 mV s−1 at different rotating speeds from 400 to 1600 rpm from 0.2 to −0.8 V. Electrochemical impedance spectrums (EIS) were measured at −0.6 V from 100 kHz to 0.01 Hz and the perturbing AC amplitude was 5 mV.

The electron transfer number (n) can be calculated by Koutecky–Levich equations as follows:

image file: c6ra09402c-t1.tif

B = 0.62nFC0(D0)2/3v−1/6
where JK is the kinetic current density; JL is the diffusion-limiting current density; J is the measured current density. B is the reciprocal of the slope; ω is angular velocity of the disk (ω = 2πN). N is the linear rotation speed. F is the Faraday constant; C0 is the saturated concentration of O2 in 0.1 mol L−1 KOH at room temperature; D0 is the diffusion coefficient of oxygen in water; v is the kinematic viscosity of the solution at room temperature. The durabilities of the N-doped graphene and Pt/C were tested by chronoamperometry (CA) measurement conducted at −0.6 V for 1800 s.

Furthermore, for capacitance performance measurement of MrGOth, cyclic voltammogram (CV) was measured at different scanning rate between −1.0 and 0.0 V in a KOH aqueous electrolyte (1 mol L−1). Meanwhile, galvanic charging–discharging (GC) tests were carried out between −0.95 and 0.0 V with different current densities. The stability of the as-prepared rGO samples was tested by CV measurements at 0.1 V s−1 with 10[thin space (1/6-em)]000 cycles.

3. Results and discussion

The powder XRD patterns of the pristine graphite, GO and the resulted rGO samples are depicted in Fig. 1(a). In comparison to the pristine graphite, a typical broad (001) peak near 12.72° with an enlarged interlayer spacing of 0.822 nm was observed from the GO pattern. This should be resulted from the incorporation of oxygen functional groups during the oxidation process. After reduction by ascorbic acid and thermal treatment, the typical peak of parent GO around 12.72° disappeared in rGO samples, indicating high exfoliation and de-oxygenation of GO. Besides, weak peaks at around 24° (002) in rGO samples, with a corresponding d002 space (0.35 nm) larger than that of the pristine graphite (0.34 nm), are attributed to the irregular stacking of wrinkled graphene nano-sheets. For the more weak intensity of the XRD pattern from MrGOth, it is indicated that MrGOth has a more irregular stacking structure than rGOA.
image file: c6ra09402c-f1.tif
Fig. 1 (a) XRD patterns of the pristine graphite, GO and the resulted rGO samples; (b) XPS spectra of GO and the resulted rGO samples; (c) XPS analysis of the C 1s peak of MrGOth; (d) XPS analysis of the N 1s peak of MrGOth; (e) different nitrogen functionalities in a graphitic plane.

To reveal the composition and the nature of carbon, nitrogen and oxygen bonds in the resulting rGO samples, XPS characterization was performed in this study and all the peaks were normalized by the intensity of C 1s peak (284.8 eV). For the XPS survey spectra from GO and as-prepared rGO samples shown in Fig. 1(b), the peaks at around 284.8 eV and 533.5 eV can be assigned to the binding energies of C 1s and O 1s respectively and it is apparent to find that the intensities of O 1s peaks of the resulting rGO samples have a great decrease comparing with that of GO sample. Furthermore, a N 1s peak at around 399.8 eV is observed in the XPS spectra of MrGOth. The calculated atomic percentages of C, O and N from GO and as-prepared rGO samples are exhibited in Table 1. The atomic ratios of carbon to oxygen for MrGOth is 19.09, which is much higher than that of GO (2.12), indicating a substantial elimination of oxygen containing groups in the thermal induced reduction, and moreover by the effect of melamine resin, 7.18% of N in introduced into the reduced graphene oxide sheet. To investigate the chemical state, the C 1s and N 1s XPS peaks of MrGOth were fitted, as given in Fig. 1(c) and (d). The C 1s peak of MrGOth can be fitted to four components, corresponding to the C[double bond, length as m-dash]C/C–C, C–N, C[double bond, length as m-dash]O and O–C[double bond, length as m-dash]O species located at 284.8, 285.9, 286.2 and 291.0 eV, respectively. The peak for O–C[double bond, length as m-dash]O (291 eV) is very weak in resulted rGO samples, indicating infrequent O–C[double bond, length as m-dash]O groups. However, some oxygen still remains in graphene, which is typical for chemically reduced graphene. Similarly, the N 1s spectra of MrGOth can be fitted into three peaks centered at 398.1, 399.7 and 401.1 eV, respectively. The peaks with lower binding energies located at 398.1 and 399.7 eV correspond to pyridinic- and pyrrolic-like nitrogen (as illustrated in Fig. 1(e)). The peak located at 401.1 eV in the high-resolution N 1s spectrum can be ascribed to “graphitic” nitrogen atoms (graphitic-center and graphitic-valley), indicating that nitrogen atoms substitute the carbon atoms in the graphene sheets and were incorporated into the carbon network.11,12

Table 1 The atomic ratios of carbon to oxygen from GO and as-prepared rGO samples
Samples Atomic percentage of C/% Atomic percentage of O/% Atomic percentage of N/%
GO 67.9 32.1 0
rGOA 91.2 8.8 0
MrGOth 88.2 4.62 7.18


The SEM morphology of the as-synthesized MrGOth sample in Fig. 2(a) shows a typical curved/wrinkled appearance as a result of the thermal expansion of GO and gasification of the melamine resin anchored at the surface of GO nanosheets with a disordered stacking. Notably, the disordered stacking could lower the total energy through van der Waals cohesion and stabilize the crumpled phase. Furthermore, TEM images also show the wrinkled structures of graphene nanosheets for MrGOth.


image file: c6ra09402c-f2.tif
Fig. 2 (a) SEM image of MrGOth, (b) and (c) TEM images of MrGOth.

The pore structure of the MrGOth sample was quantitatively analyzed by the nitrogen adsorption and desorption isotherms as shown in Fig. 3(a). According to the IUPAC classification, the nitrogen adsorption isotherms of the as-prepared rGO was characterized as type IV with hysteresis loops, implying the presence of a large number of mesopores due to the formation of wrinkled and folded few-layered graphene sheets.14 The determined BET surface area of the as-synthesized rGO sample is 863.5 m2 g−1 and the pore volume is up to 1.986 cm3 g−1. According to Fig. 3(b), most of the pores focus on that with diameters sub 10 nm and the average pore diameter is about 3.4 nm in the MrGOth “fluffy” powder. The specific surface area value is lower than the theoretical surface area of 2630 m2 g−1, indicating an average overlap and stacking of ca. 3–4 layers of graphene sheets in the obtained rGO.


image file: c6ra09402c-f3.tif
Fig. 3 (a) The nitrogen adsorption and desorption isotherms of the as-prepared MrGOth and (b) surface area distribution with diameter.

All the above results demonstrate the unique porous, crumpled structure of MrGOth and high nitrogen doping contents. In the thermal treatment of the surface functional GO, the melamine resin could react with GO matrix on the defects and vacancies. For melamine resin anchored on both sides of GO ultrathin sheet, it enabled the nitrogen doping reaction to proceed on both sides of GO, leading to uniform and high content nitrogen doping in graphene.13 Therefore, it differs with the nitrogen doped graphene using ammonia gas as the nitrogen source, in which the reaction mostly occurred on the exposed surface of carbon nanostructures, resulting in inhomogeneous and low nitrogen doping content, of the order of 2–4 at%.15–18

The electrochemical activity for oxygen reduction reaction of the resulting rGO samples was investigated by linear scanning voltammogram on RDE with a rotation from 400–1600 rpm. According to the LSV results shown in Fig. 4(a) and S2, it is apparent to see that both rGOA and MrGOth hold electrochemical catalytic activity to oxygen reduction reaction. Furthermore, LSV shows that limiting current density increases with higher rotation speeds as shown in Fig. 4(a) and (b) and S2. The phenomenon can be explained by the shortened diffusion distance at high speeds, which is in accordance with other studies.18–22 The ORR performance of MrGOth in the diffusion and kinetically limited regions was evaluated using Koutecky–Levich (K–L) plots as shown in Fig. 4(c) and the K–L plots of RDE curves show a linear relationship between 1/j and ω1/2 from −0.5 to −0.7 V. According to Koutecky–Levich equation, the number of electrons transferred (n) in the ORR process can be calculated and the n value was increased as the applied potential rose. As shown in Fig. 4(d), at the potential of −0.7 V, n of MrGOth reaches to 3.4, and for 20% Pt/C catalyst is 3.9, indicating that MrGOth follows an almost four-electron transfer pathway and has better ORR activity than rGOA. At the 1600 rpm rotating speed, the LSV curves of rGOA, MrGOth and Pt/C catalysts were shown in Fig. 4(e) and it is obvious to see that the current density follow the order of rGOA < MrGOth < Pt/C. The onset potential of MrGOth is about −0.12 V, which is only slightly more negative compared to that of Pt/C (−0.08 V). As shown in Fig. 4(f), no matter at the potential of −0.7 V, or at −0.6 V, even at −0.5 V, the current densities of MrGOth are up to 80% of the current density of Pt/C. Fig. 4(g) reveals that the ORR electrochemical activity of MrGOth is higher than most of the reported N-doped carbon materials.23–32 Unlike MrGOth, rGOA exhibited a two-step two-electron process for oxygen reduction with the onset potential of about ∼0.26 V. The catalytic current density was found to be 52% less than that of MrGOth at the potential of −0.8 V. These results indicate that the doping of N atoms in the carbon ring plays a key role in improving the catalytic activity. To further manifest the catalytic activity improvement of the MrGOth, EIS was measured on a rotation disk electrode at −0.6 V with a rotation of 1600 rpm. Fig. 4(h) represents the Nyquist plots of different catalysts and the data were analyzed by the electric equivalent circuit shown in Fig. 4(i). In the equivalent electric circuit, Rs is the sum of resistance of electrolyte, electrode material and the contact resistance at the interface of the active material/current collector; Q1 is a constant phase element, and Y1 and n1 are the parameters of Q1; C1, which represents the double layer capacitance, can be calculated by C1 = [Y1 × R1^(1 − n1)]^(1/n); R1 is the resistance of the double layer; Q2 represents the property of intermediate adlayer and R2 is charge transfer resistance. The values of Rs, Q1, R1, Q2 and R2 were calculated from the CNLS fitting of the experimental impedance spectra and their resulting values are listed in Table 2. According to the fitting results of C1 and R1, it can be concluded that rGOA has the weakest ability to absorb O2, which induced the largest resistance in O2 absorbing process and MrGOth has a similar ability to absorb O2 with Pt/C catalyst. The results of C2 indicate that both MrGOth and Pt/C catalysts have fewer amounts of intermediates than rGOA, due to their nearly direct 4 electron transfer in O2 reduction process. Moreover, the charge transfer resistance has a great decrease from rGOA to MrGOth. Therefore, all these EIS results manifest that MrGOth has a better electrocatalytic activity than rGOA by N-doping. Furthermore, the stabilities of MrGOth and Pt/C were tested by chronoamperometry (CA) measurements conducted at −0.6 V for 1800 s. All the data collected were normalized by the initial current density and shown in Fig. 4(j). It is apparent to see that the current density of Pt/C decreased to 88% of the initial current density after 1800 s test, which means the deterioration of the electrocatalytic activity. But for MrGOth, the current density has an increase before 400 s and then a decrease. After 1800 s test, the current density is still 110% of the initial one. Therefore, it manifests that MrGOth has a better durability than Pt/C.


image file: c6ra09402c-f4.tif
Fig. 4 (a) ORR polarization curves (at different rotation speeds) of MrGOth; (b) ORR polarization curves (at different rotation speeds) of Pt/C; (c) Koutechky–Levich plots at different potentials of MrGOth; (d) comparison of electron-transfer number from different catalysts; (e) ORR polarization curves of different catalysts at 1600 rpm; (f) comparison of current densities at different potentials from different catalysts; (g) current density comparison with results previous reported;23–32 (h) Nyquist plots of different catalysts at −0.6 V; (i) electric equivalent circuit; (j) proportion to the initial current density.
Table 2 Fitting results of EIS from different catalysts
  Rs (Ω) C1 (F) R1 (Ω) C2 (F) R2 (Ω)
rGOA 3.645 6.228 × 10−6 782 7.71952 × 10−5 8044
MrGOth 3.868 9.16284 × 10−6 88.54 4.30614 × 10−5 7170
Pt/C 3.513 9.58932 × 10−6 189.4 3.99385 × 10−5 5619


To have a deep insight in ORR catalytic activity improvement from rGOA to MrGOth by N-doping, the oxygen adsorptions on graphene and N-doped graphene were studied based on Density Functional Theory (DFT) by using Materials studio. According to results shown in Fig. 5, it is apparent to know that for graphene, the oxygen molecule is physically adsorbed on the center position above the C6-ring, which is 2.824 Å from the graphene sheet. The length of the chemical bond in oxygen molecule is about 1.263 Å, which is longer than that of oxygen molecule in vacuum (1.235 Å). Furthermore, for N-doped graphene, the adsorption position of oxygen molecule is changed by the effect of the doped nitrogen atom. One of the atoms in O2 is located on the top of carbon atom neighbored nitrogen atom and it is 2.461 Å apart from this carbon atom. The other oxygen atom is above the center of C5N-ring and 2.775 Å apart from. The bond length in oxygen molecule enlarges to about 1.280 Å. It indicates that the doped nitrogen atom in graphene could increase the bond length of the adsorbed O2 and decrease the bond energy.33,34 Therefore, O2 absorbed on N-doped graphene is more active than that on pure graphene.


image file: c6ra09402c-f5.tif
Fig. 5 O2 adsorbed on pure graphene (a) and on N-doped graphene (b).

In order to investigate the effects of N-doping on supercapacitance of reduced graphene oxide, cyclic voltammetry (CV) and galvanostatic charge–discharge tests were carried out using a three electrode system. Fig. 6(a) shows the CV curves of rGOA and MrGOth at the scan rate of 0.05 V s−1. The rectangular CV curves with a pair of broaden peaks were observed, which indicate the pseudo-capacitance was introduced into the double layer capacitance. Additionally, MrGOth electrode exhibits much higher current density than the electrode of rGOA, indicating increased capacitance due to N-doping. Cyclic voltammetry of rGOA and MrGOth at various scan rates are shown in Fig. S3. As the scan rate increases, the CV curves of rGOA have more obvious deviations from rectangularity than that of MrGOth. It indicates MrGOth has a more excellent rate capability than rGOA. The improved capacitance of MrGOth can also be observed from the galvanostatic charge–discharge curves at a current density of 2 A g−1 as shown in Fig. 6(b). The galvanostatic charge–discharge curves of as-prepared rGO samples at various current densities are shown in Fig. S4. All these charge–discharge curves at different current densities show linear responses and good symmetry, demonstrating the excellent double layer capacitive behavior. Fig. 6(c) shows the dependence of the specific capacitance of rGOA and MrGOth on the discharge current densities. It is clearly to conclude that MrGOth has a better rate capability than rGOA. This may be attributed to the synergistic effects of high specific surface area and N-doping of MrGOth. As Fig. 6(c) shown, the specific capacitance of MrGOth is 163 F g−1 at the current density of 1 A g−1. Due to the synergistic effects of high surface area and N-doping, the specific capacitance is still as high as 108 F g−1 even at the current density of 100 A g−1. When the current density increases to 200 A g−1, the specific capacitance also remains to 78 F g−1, indicating the MrGOth shows the best rate capability to our known.35–37 The electrochemical stability of MrGOth electrode for supercapacitor application was also evaluated using the cyclic voltammetry at a scan rate of 0.1 V s−1. As shown in Fig. 6(d), after 10[thin space (1/6-em)]000 cycles there is no significant change in the shape of CV curves. The capacitance retention of MrGOth is 92.68% after 10[thin space (1/6-em)]000 cycles as shown in Fig. 6(e). Therefore, it demonstrates that MrGOth has excellent electrochemical stability and a high degree of reversibility.


image file: c6ra09402c-f6.tif
Fig. 6 (a) CVs of as-prepared samples at the scan rate of 0.05 V s−1; (b) galvanostatic charge–discharge curves of as-prepared samples at charge current density of 2 A g−1; (c) specific capacitance of as-prepared samples at different discharge current densities; (d) CV cycling performance of MrGOth; (e) CV cycling stability at a scan rate of 0.1 V s−1.

The above results clearly reveal that MrGOth exhibits significantly improved electrochemical capacitance compared with rGOA. It should be noted that MrGOth has a high nitrogen doping content (the ratio of C to N is 12.28) with apparent crumpled structures which can help it achieve high capacitance, excellent rate capability and cycle performance. As reported by previous works, there are four reasons for the improved supercapacitance: firstly, pyridinic N and pyrrolic N are assumed to be the main configurations constructing to the faradic reaction based pseudo-capacitance. It is also reported that the presence of quaternary N can enhance the conductivity of the materials, which is beneficial for rate capability of the supercapacitor performance.34,38,39 Additionally, the heterogeneous nitrogen atom, which is introduced into carbonaceous structures, could change the electron distribution of the materials, which can further enhance the wettability between electrode materials and electrolyte, leading to a significant increase of active surface area accessible to the electrolyte.34 Furthermore, the remaining oxygen-containing surface functional groups such as hydroxyl and carbonyl groups can provide pseudo-capacitance as a complement to the capacitance. Finally, the high temperature thermal-treatment can increase the conductivity of carbon-based materials and give a highly crumpled structure to the reduced graphene oxide. Therefore, it can provide a high accessible specific surface area for adsorbing ions and accelerating electron transfer. Therefore, it resulted in an excellent rate capability of MrGOth compared with many other reported samples.35–37

4. Conclusions

N-Doped porous reduced graphene oxide was successfully obtained by a melamine resin surface finishing and thermal treatment method. The physical characterizations demonstrated that the nitrogen atoms were introduced into the reduced graphene oxide by this method. Based on the high specific surface area, high porosity, and doped hetero-atoms, the as-prepared MrGOth shows excellent electrocatalytic activity towards the ORR and high capacitance compared with reduced graphene oxide sample. Therefore, this work demonstrates a good example of both electrocatalytic activity and capacitance properties of N-doped rGO prepared by surface decorating of GO and followed heat treatment method, which opens the door for creating functional graphene materials with highly promising applications in high-performance renewable energy conversion and storage devices.

Acknowledgements

The authors are grateful for the financial support by One Hundred Talent Program of Chinese Academy of Sciences, Chinese National Programs for High Technology Research and Development (2014AA06A513), as well as by the NSFC (51302264, 51404230) of China.

References

  1. R. M. John and S. Patrice, Science, 2008, 321, 651 CrossRef PubMed.
  2. D. Liu, B. B. Garcia, Q. Zhang, Q. Guo, Y. Zhang, S. Sepehri and G. Cao, Adv. Funct. Mater., 2009, 19, 1015 CrossRef CAS.
  3. L. Estevez, R. Dua, N. Bhandari, A. Ramanujapuram, P. Wang and E. P. Giannelis, Energy Environ. Sci., 2013, 6, 1785 CAS.
  4. H. Pang, C. Z. Wei, X. X. Li, G. C. Li, Y. H. Ma, S. J. Li, J. Chen and J. S. Zhang, Sci. Rep., 2014, 4, 3577 Search PubMed.
  5. E. Frackowiak, Phys. Chem. Chem. Phys., 2007, 9, 1774 RSC.
  6. S. J. Amirfakhri, D. Binny, J. L. Meunier and D. Berk, J. Power Sources, 2014, 257, 356 CrossRef CAS.
  7. J. W. Lee, J. M. Ko and J. D. Kim, Electrochim. Acta, 2012, 85, 459 CrossRef CAS.
  8. M. W. Chung, C. H. Choi, S. Y. Lee and S. I. Woo, Nano Energy, 2015, 11, 526 CrossRef CAS.
  9. D. Kubo, K. Tadanaga, A. Hayashi and M. Tatsumisago, J. Mater. Chem. A, 2013, 1, 6804 CAS.
  10. H. Y. Zhao, C. H. Sun, Z. Jin, D. W. Wang, X. C. Yan, Z. G. Chen, G. S. Zhu and X. D. Yao, J. Mater. Chem. A, 2015, 3, 11736 CAS.
  11. D. S. Geng, N. Ding, T. S. A. Hor, Z. L. Liu, X. L. Sun and Y. Zong, J. Mater. Chem. A, 2015, 3, 1795 CAS.
  12. Q. Li, R. G. Cao, J. Cho and G. Wu, Adv. Energy Mater., 2014, 4, 1301415 Search PubMed.
  13. J. H. Lee, N. Park, B. G. Kim, D. S. Jung, K. Im, J. Kur and J. W. Choi, ACS Nano, 2013, 7, 9366 CrossRef CAS PubMed.
  14. Z. J. Jia, J. Wang, Y. Wang, T. Qi and C. M. Li, Adv. Electron. Mater., 2015, 1, 11 Search PubMed.
  15. X. D. Huang, Y. F. Zhao, Z. M. Ao and G. X. Wang, Sci. Rep., 2014, 4, 7557 CrossRef CAS PubMed.
  16. P. Chen, J. J. Yang, S. S. Li, Z. Wang, T. Y. Xiao, Y. H. Qian and S. H. Yu, Nano Energy, 2013, 2, 249 CrossRef CAS.
  17. S. Y. Gao, Y. L. Chen, H. Fan, X. J. Wei, C. G. Hu, H. X. Luo and L. T. Qu, J. Mater. Chem. A, 2014, 2, 3317 CAS.
  18. X. J. Bo, M. Li, C. Han, Y. F. Zhang, A. Nsabimana and L. P. Guo, J. Mater. Chem. A, 2015, 3, 1058 CAS.
  19. Y. Chen, J. Li, T. Mei, X. G. Hu, D. W. Liu, J. C. Wang, M. Hao, J. H. Li, J. Y. Wang and X. B. Wang, J. Mater. Chem. A, 2014, 2, 20714 CAS.
  20. K. C. Poon, D. C. L. Tan, T. D. T. Vo, B. Khezri, H. B. Su, R. D. Webster and H. Sato, J. Am. Chem. Soc., 2014, 136, 5217 CrossRef CAS PubMed.
  21. X. Y. Yan, X. L. Tong, Y. F. Zhang, X. D. Han, Y. Y. Wang, G. Q. Jin, Y. Qin and X. Y. Guo, Chem. Commun., 2012, 48, 1892 RSC.
  22. M. K. Liu, Y. F. Song, S. X. He, W. W. Tjiu, J. S. Pan, Y. Y. Xia and T. X. Liu, ACS Appl. Mater. Interfaces, 2014, 6, 4214 CAS.
  23. C. Y. He, Z. S. Li, M. L. Cai, M. Cai, J. Q. Wang, Z. Q. Tian, X. Zhang and P. K. Shen, J. Mater. Chem. A, 2013, 1, 1401 CAS.
  24. D. Higgins, Z. Chen, D. U. Lee and Z. W. Chen, J. Mater. Chem. A, 2013, 1, 2639 CAS.
  25. L. F. Lai, J. R. Potts, D. Zhan, L. Wang, C. K. Poh, C. H. Tang, H. Gong, Z. X. Shen, J. Y. Lin and R. S. Ruoff, Energy Environ. Sci., 2012, 5, 7936–7942 CAS.
  26. H. P. Cong, P. Wang, M. Gong and S. H. Yu, Nano Energy, 2014, 3, 55 CrossRef CAS.
  27. N. Gavrilov, I. A. Pasti, M. Mitric, J. Travas-Sejdic, G. Ciric-Marjanovic and S. V. Mentus, J. Power Sources, 2012, 220, 306 CrossRef CAS.
  28. X. J. Zhao, H. Y. Zhao, T. T. Zhang, X. C. Yan, Y. Yuan, H. M. Zhang, H. J. Zhao, D. M. Zhang, G. S. Zhu and X. D. Yao, J. Mater. Chem. A, 2014, 2, 11666 CAS.
  29. L. F. Wu, H. B. Feng, M. J. Liu, K. X. Zhang and J. H. Li, Nanoscale, 2015, 5, 10839 RSC.
  30. S. Y. Gao, X. J. Wei, H. Fan, L. Y. Li, K. R. Geng and J. J. Wang, Nano Energy, 2015, 13, 218 CrossRef.
  31. Y. W. Zhang, J. Ge, L. Wang, D. D. Wang, F. Ding, X. M. Tao and W. Chen, Sci. Rep., 2013, 3, 2771 Search PubMed.
  32. H. J. Yan, B. Xu, S. Q. Shi and C. Y. Ouyang, J. Appl. Phys., 2012, 112, 104316 CrossRef.
  33. X. D. Ren, J. Z. Zhu, F. M. Du, J. J. Liu and W. Q. Zhang, J. Phys. Chem. C, 2014, 118, 22412 CAS.
  34. Y. Q. Zou, I. A. Kinloch and R. A. W. Dryfe, J. Mater. Chem. A, 2014, 2, 19495 CAS.
  35. X. L. Fang, J. Zang, X. L. Wang, M. S. Zheng and N. F. Zheng, J. Mater. Chem. A, 2014, 2, 6191 CAS.
  36. P. F. Zhang, Z. A. Qiao, Z. Y. Zhang, S. Wan and S. Dai, J. Mater. Chem. A, 2014, 2, 12262 CAS.
  37. P. Hao, Z. H. Zhao, Y. H. Leng, J. Tian, Y. H. Sang, R. I. Boughton, C. P. Wong, H. Liu and B. Yang, Nano Energy, 2015, 15, 9 CrossRef CAS.
  38. Y. F. Zhao, W. Ran, J. He, Y. Z. Huang, Z. F. Liu, W. Liu, Y. F. Tang, L. Zhang, D. W. Gao and F. M. Gao, Small, 2015, 11, 1310 CrossRef CAS PubMed.
  39. H. N. Ma, J. He, D. B. Xiong, J. S. Wu, Q. Q. Li, V. Dravid and Y. F. Zhao, ACS Appl. Mater. Interfaces, 2016, 8, 1992 CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra09402c

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.