Needle-like MnO2/activated carbon nanocomposites derived from human hair as versatile electrode materials for supercapacitors

Dian Denga, Byoung-Suhk Kimb, Mayakrishnan Gopiraman*a and Ick Soo Kim*a
aNano Fusion Technology Research Lab, Institute for Fiber Engineering (IFES), Division of Frontier Fibers, Interdisciplinary Cluster for Cutting Edge Research (ICCER), National University Corporation, Shinshu University, Ueda, Nagano 386-8567, Japan. E-mail: kim@shinshu-u.ac.jp; gopiramannitt@gmail.com
bDepartment of Organic Materials & Fiber Engineering, Chonbuk National University, 567 Baekje-daero, Deokjin-gu, Jeonju-si, Jeollabuk-do 561-756, Republic of Korea

Received 18th August 2015 , Accepted 17th September 2015

First published on 17th September 2015


Abstract

Bio-waste material, human hair, was used as a source to prepare activated carbon (ACs) by a simple NaOH activation method. MnO2 with a needle like structure was successfully decorated on ACs by a wet impregnation method. TEM images confirmed the needle-like morphology of MnO2 on ACs. The chemical state of Mn in MnO2/ACs was found to be +4, as confirmed by XPS analysis. Other physicochemical properties of MnO2/ACs were investigated by means of SEM-EDS, XRD, Raman and XPS analyses. After complete characterization, the resultant nanocomposites with different MnO2-loading [MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y)] were employed as electrode materials for supercapacitors. Interestingly, the MnO2/ACs nanocomposites showed an excellent capacitance in three different electrolytes (1.0 M H2SO4, 1.0 M KOH and 1.0 M Na2SO4). The MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) achieved a maximum capacitance of 410 F g−1, 345 F g−1 and 291 F g−1 in 1.0 M H2SO4, 1.0 KOH and 1.0 Na2SO4, respectively. To the best of our knowledge, this is the first MnO2-based carbon nanocomposite to show good capacitance performance in three different kinds of electrolytes. Moreover, the MnO2/ACs showed a capacitance of ∼300 F g−1 even after 500 cycles in 1.0 M H2SO4.


1. Introduction

Supercapacitors have recently gained huge attention due to their advantages such as longer cycle life, short charging time and higher power density compared to conventional electrical double-layer capacitors.1 Several carbon-based electrode materials such as activated carbon (ACs), graphene, carbon nanotubes (CNTs), fullerenes, and carbon fibers are reported as superior supercapacitor electrode materials.2,3 Among them, due to low cost, very high specific surface area and pore properties, ACs are often preferable over other carbon materials.4 Recently, Qian et al.,5 prepared carbon flakes from human hair and employed as electrode materials for supercapacitors. They found that the carbon flakes have a very high specific capacitance of 340 F g−1 in 6.0 M KOH at a current density of 1 A g−1 and good stability over 20[thin space (1/6-em)]000 cycles. Alike, Ma and co-workers6 used cotton fiber as carbon source to prepare porous activated carbon (PACs). As an electrode material, the prepared PACs showed a high specific capacitance of 239 F g−1 at 0.5 A g−1 current density and good rate capability in 2.0 M KOH aqueous electrolyte. Alike, ACs derived from plant leaves exhibited a specific capacitance of 400 F g−1 and an energy density of 55 W h Kg−1 in 1.0 M H2SO4.7 Most of the ACs derived from biomass is performed as electrode materials either in acid or base electrolyte.8,9 However, the drawback of these ACs is its poor capacitance performance in neutral electrolytes such as Na2SO4 and K2SO4; only a very few reports exist up to now on the capacitance performance of ACs in neutral electrolytes.10 Hence, the development of a versatile supercapacitor electrode material (which can perform in various electrolytes such as acid, base and neutral media) is a challenging task.

Modification or functionalization of porous ACs with metal oxides such as MnOx, RuOx, NiOx and CuOx is one of the promising approaches to achieve superior electrode performance.11,12 Among them, manganese oxide (MnO2) is often preferred as an electrode material for supercapacitors due the low cost, high theoretical surface area and non-toxic.13 Moreover, the MnO2 can exhibit larger capacitance values and energy density due to its fast surface redox reactions. The MnO2 has showed a maximum specific capacitance of 1370 F g−1.14 However, the pseudocapacitive reaction of MnO2 is known to be a surface reaction that can occur only on the surface. Hence, the MnO2 is often combined with conductive materials such as graphene, CNTs and carbon fibers to improve its performance.15 Yan et al.,16 prepared MnO2/graphene and employed as supercapacitor electrode material. The MnO2/graphene demonstrated a very high capacitance of 310 F g−1. MnO2 nanorods were supported on ACs and employed as electrode material for supercapacitor.17 It showed a moderate capacitance of 165 F g−1 in Na2SO4 electrolyte. In case of MnO2/activated carbon composites, the performance is mainly dependent on five factors:18,19 (a) surface area of the composites, (b) structure of the MnO2 and ACs, (c) conductivity of the materials, (d) large pore volume and pore size, and (e) presence of heteroatom such as O, N and S. Recently, human hair has been used to prepare ACs with very high surface area and excellent pore structure. In addition, the ACs found to consist of various heteroatoms such as O, N and S, and has exhibited very high specific capacitance in 6 M KOH electrolyte.5 We believe that the functionalization of MnO2 with unique structure on ACs (derived from human hair) would show superior capacitance performance in various electrolytes.

Herein, we prepared needle-like MnO2/ACs nanocomposites with different MnO2 loading by a simple wet impregnation method. The ACs was derived from human hair by NaOH activation. The prepared composite materials were characterized by various spectroscopic and microscopic methods such as HR-TEM, SEM-EDS, XRD, Raman and XPS analyses. After complete characterization, the resultant nanocomposites MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y)] were employed as electrode materials for supercapacitor in three different electrolytes, namely, 1.0 M H2SO4, 1.0 M KOH and 1.0 M Na2SO4. Conductivity and cycle life of the nanocomposites were also studied.

2. Experimental

2.1 Materials

Human hair was collected from students of Shinshu University, Japan. Nafion®perfluorinated resin solution (5 wt%), 2-propanol, potassium permanganate (KMnO4) were purchased from Sigma Aldrich. Sodium sulfate (Na2SO4), potassium hydroxide (KOH), sulfuric acid (H2SO4), hydrochloric acid (HCl) and sodium hydroxide (NaOH) were purchased from Wako Pure Chemicals, Japan. All chemicals were used without further purification.

2.2 Preparation of ACs

The ACs was prepared from human hair according to the previously reported procedure.5 In a typical experiment, 1.0 g of human hair was washed well and dried in air. At first, the human hair was stabilized under air atmosphere at 300 °C (heating rate of 1°C min−1) for 1 h. Subsequently, the stabilized human hair was mixed with NaOH (human hair/NaOH = 1[thin space (1/6-em)]:[thin space (1/6-em)]2 ratio) using mortar and pestle until the homogeneous mixture was obtained. Finally, the mixture was calcinated in the muffle furnace under N2 atmosphere at 700 °C (heating rate of 5°C min−1) for 3 h. After calcination, the resultant solid was washed well with 1 M HCl to remove the NaOH.

2.3 Synthesis of MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) nanocomposites

In a typical synthesis procedure, 25 mg of ACs was dispersed in 10 mL distilled water followed by sonication for 30 min. To the above mixture, KMnO4 solution was added drop by drop under vigorous stirring condition. After the addition of KMnO4, the mixture was stirred at 60 °C for 24 h and then vacuum filtered to obtain the MnO2/ACs. ACs and KMnO4 with different weight ratios (ACs/KMnO4 = 1[thin space (1/6-em)]:[thin space (1/6-em)]2, 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 2[thin space (1/6-em)]:[thin space (1/6-em)]1, 4[thin space (1/6-em)]:[thin space (1/6-em)]1, 8[thin space (1/6-em)]:[thin space (1/6-em)]1, 10[thin space (1/6-em)]:[thin space (1/6-em)]1, 12[thin space (1/6-em)]:[thin space (1/6-em)]1, 14[thin space (1/6-em)]:[thin space (1/6-em)]1 and 16[thin space (1/6-em)]:[thin space (1/6-em)]1) were prepared. Here, the nanocomposites are denoted as MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y), where X is the weight ratio of ACs and Y represents the weight ratio of KMnO4. Finally, the MnO2/ACs was washed well with water and air dried. The resultant composites were characterized in detail and used for electrochemical studies.

2.4 Characterization

The surface morphology of the resultant MnO2/ACs composites was studied by field emission scanning electron microscopy (FE-SEM; S-5000, Hitachi) and transmission electron microscopy (TEM; 2010FasTEM, JEOL, Japan). Energy dispersive X-ray spectra (EDS) were performed on JEOL JSM-5900. X-ray photoelectron spectroscopy (XPS; AXIS-ULTRADLD, Shimadzu) and Raman spectrometer (Hololab 5000, Kaiser Optical Systems, Inc., Ann Arbor, MI, USA) were recorded to study the chemical state of the atoms presented in the MnO2/ACs. X-ray diffraction (XRD) experiment was performed at 27 °C using a Rotaflex RTP300 (Rigaku Co., Japan) diffractometer at 50 kV and 200 mA with a scan speed of 2°/min. Nickel-filtered Cu Kα radiation (5 < 2θ < 70°) was used for XRD measurements. For Raman analysis, Ar laser was operated at 532 nm with a Kaiser holographic edge filter. The specific surface area of samples was determined using the Brunauer–Emmett–Teller (BET) method (BELSORP-max, BEL Japan, Inc.). The electrochemical studies were carried out in Versastat-4 potentiostat at a scan rate ranging from 5 to 100 mV s−1.

2.5 Electrochemical measurements

The electrochemical measurements were carried out in three different electrolytes (1.0 M H2SO4, 1.0 M KOH and 1.0 M Na2SO4) at 27 °C using Versastat 4 electrochemical station. The sweep potential range was adjusted from −0.2 to 0.8 V [vs. Ag/AgCl] in an electrochemical cell with three-electrode system: platinum wire, Ag/AgCl and prepared carbon nanocomposites were used as counter electrode, reference electrode and working electrode, respectively. The electrochemical impedance spectroscopy (EIS) was recorded in the frequency range of 0.01 Hz to 100 kHz with a potential amplitude of 10 mV. The working electrode was prepared as follows; 2 mg of nanocomposite, 20 μL of Nafion solution (5 wt%) and 400 μL of isopropanol were mixed and ultrasonicated at 27 °C for 2 h. The 30 μL of the prepared slurry was deposited on the active area of the glassy carbon electrode and dried at 80 °C for 30 minutes to evaporate the isopropanol.

3. Results and discussion

The MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) nanocomposites with different MnO2 loading were prepared and the physicochemical characteristics were investigated in detail. The representative data of SEM-EDS, BET, Raman, XRD and XPS obtained for pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) are shown and discussed. Data of other samples [MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y)] are given in the ESI (Fig. S1–S4). The BET pore properties and EDS elemental compositions of ACs and nanocomposites are presented Table 1. Fig. 1 represents the EDS weight percentage of elements (C, O, S and Mn) presented in MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) nanocomposites. As expected, the pure ACs showed 0% of Mn with C (85.69%), O (12.50%) and S (2.06%). Although the presence of N was detected by XPS analysis, the EDS result didn't show any peak corresponds to N which may be due to the presence of a trace amount of N. In case of MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) nanocomposites, the wt% of Mn was gradually increased with increasing KMnO4 ratio (Table 1). Fig. 2 shows the representative SEM image, EDS spectrum and its corresponding elemental mappings (C, Mn and O) of MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12). The weight percentage of Mn, O, S and C in MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) nanocomposites was found to be 4.72, 12.38, 3.17 and 79.73, respectively. In addition, homogeneous dispersion of MnO2 was confirmed by the elemental mapping of Mn [Fig. 2(iii)]. For more details, refer Fig. S1 in ESI. Fig. 3 represents nitrogen adsorption–desorption isotherms of pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12). The pure ACs showed the specific surface area of 1318.4 m2 g−1. Alike, the BET surface areas of 1597.4 m2 g−1 was observed for MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12). In addition, the MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) has total pore volume and average pore diameter of 1.02 cm3 g−1 and 2.5542 nm, respectively (Table 1). This high surface area can offer a sufficient electrode/electrolyte interface for ion or charge accumulation. Moreover, the presence of high volume of pores (1.02 cm3 g−1) in the materials can assist for the rapid diffusion of ions and results in a remarkably improved rate performance of electrode materials.20
Table 1 BET pore characteristics and EDS elemental compositions of ACs and its MnO2-composites
Samples BET analysis EDS (wt%)
SBET (m2 g−1) Daver (nm) Vpore (cm3 g−1) C O S Mn
Pure ACs 1318.4 2.40 0.79 85.69 12.05 2.06
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2) 106.8 7.14 0.19 42.12 21.78 1.94 34.17
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]1) 636.0 3.41 0.39 45.51 21.58 2.48 30.43
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2) 600.2 2.60 0.54 60.62 23.52 2.37 13.49
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]4) 1355.7 2.38 0.81 72.44 15.38 2.07 10.11
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]8) 1519.3 2.51 0.95 77.33 15.65 2.46 4.57
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]10) 1560.9 2.35 0.92 77.87 14.23 2.72 5.18
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) 1597.4 2.55 1.02 79.73 12.38 3.17 4.72
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]14) 1705.3 2.68 1.14 81.73 12.12 2.28 3.87
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]16) 1641.4 2.00 0.82 85.12 10.16 2.53 2.19



image file: c5ra16624a-f1.tif
Fig. 1 EDS elemental weight percentage of (a) pure ACs, (b) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2), (c) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]1), (d) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2), (e) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]4), (f) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]8), (g) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]10), (h) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12), (i) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]14), and (j) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]16).

image file: c5ra16624a-f2.tif
Fig. 2 (i) SEM image of MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) and corresponding elemental mapping of (ii) C, (iii) Mn and (iv) O. (v) EDS spectrum of MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12).

image file: c5ra16624a-f3.tif
Fig. 3 HR-TEM images of (i) ACs, (ii and iii) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2) and (iv–vi) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) [images (iii) and (vi) are magnified images of MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2) and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12), respectively].

The HR-TEM images were taken to study the microstructure of the ACs [Fig. 3(i)], MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2) [Fig. 3(ii and iii)] and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) [Fig. 3(iv–vi)]. As seen from HR-TEM images of pure ACs in Fig. 3(i), the architecture of ACs with two-dimensional (2D) nanosheet morphology and the average thickness of the sheets were calculated to be ∼65 nm. In addition, meso/micropores channels can be clearly seen on the surface of the nanosheets. In Fig. 3, the MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) composite showed that the needle like MnO2 was homogeneously anchored on the 2D nanosheet. The average diameter and average length of the MnO2 needles are found to be ∼14 nm and ∼110 nm, respectively. Interestingly, the average thickness of the nanosheets was dramatically decreased from 65 to 41 nm compared to pure ACs. In general, the inter-layers of carbon stacked by van der Waals forces have interaction energy of ∼2 eV nm−2 and typically a very weak ∼300 nN lm−2 magnitude of force is required to break this energy.21,22 In the present study, the exfoliation of the carbon sheets might have achieved during the preparation of nanocomposite (stirring and sonication process). In addition, the formed MnO2 may be stabilized or prevented the carbon nanosheets from the further regeneration. The HR-TEM result agrees well with the BET results. Moreover, it was found that the loading of KMnO4 has greater influence on the structure of the MnO2. At higher loading of MnO2, the surface morphology of the nanocomposites was quite different and no needle-like MnO2 structures were observed on the carbon nanosheets. Instead, the MnO2 was fully covered the carbon sheets at higher loading of KMnO4. The needle-like nanostructure of MnO2 can help to an easy ionic charge transport at electrode/electrolyte interfaces.23

XRD and Raman spectra were recorded to further investigate the graphitic structure [ordered and disordered (defect sites) nature] of pure ACs and nanocomposites (Fig. 4, 5, S3 and S4). Fig. 4 shows the XRD patterns of pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12). Two characteristic peaks were seen at around 2θ = 22.3° corresponding to (002) and 2θ = 43.8° (a weak peak) corresponding to (101) plane.24 The XRD peak at around 2θ = 22.3° (002) attributed to a well-defined graphitic stacking. Alike, the weak peak at 2θ = 43.8° supports to the higher degree of interlayer condensation of carbon. Moreover, the clear observation of these XRD peaks indicating that the pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) consist of small domains of prearranged graphene sheets. These results support that the samples are highly conductive in nature.25 Raman spectra of both ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) showed two characteristic bands at 1325 (D band) and 1597 cm−1 (G band).26 The G-band was invented from the in-plane vibration of sp2 carbon atoms, which attributed to the graphitic carbon. The D-band line was related to the amount of disorder, indicating the presence defect sites in the carbon network. Generally, the D/G ratio of band intensities is often used to study defects concentration in carbon.27 At first, the D band and G band were fitted as the sum of a Gaussian function and then the IG/ID values were calculated (Fig. 5). The IG/ID values of ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) were calculated to be 1.0167 and 1.0601, respectively. The Raman IG/ID values confirm a highly graphitized ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12), offering an excellent electric conductivity, which is consistent with the XRD results.28


image file: c5ra16624a-f4.tif
Fig. 4 XRD spectra of ACs (a) and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) (b).

image file: c5ra16624a-f5.tif
Fig. 5 Raman spectra of ACs (a) and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) (b).

Further, the XPS spectra were recorded for pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12). Fig. 6 shows the high-resolution XPS spectra of C 1s, N 1s, O 1s, S 2p, and Mn 2p. It was confirmed that the chemical compositions of pure ACs and nanocomposites are consist of C, N, O and S. The narrow and intense C 1s XPS peak was observed for both ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12), indicating an improved degree of graphitic order. Alike, the O 1s peak located at B.E. = 533.5 eV shows the presence of oxygen atoms. In fact the oxygen functional groups can enhance wettability of the carbon materials.29 A strong and broad peak at B.E. = 400 eV (N 1s XPS peak) was observed which reveal the presence of N species in the form of pyridinic, pyrrolic, N-oxide and quaternary nitrogen.30 Similarly, the S appears at approximately B.E. = 165 eV. In fact the presence of heteroatom such as S, N and O can contribute greatly to the pseudocapacitance. Two obvious peaks at B.E. = 642.0 eV and B.E. = 653.8 eV are observed in the Mn 2p XPS spectrum, corresponding to Mn 2p3/2 and Mn 2p1/2 peak, respectively. The B.E. values of Mn 2p peaks confirmed the +4 oxidation state of Mn in MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12).31


image file: c5ra16624a-f6.tif
Fig. 6 (i) Survey XPS spectra, (ii) S 2s peaks, (iii) C 1s peaks, (iv) N 1s peaks and, (v) O 1s peaks of ACs (a) and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12), and (vi) XPS Mn 2p peak of MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12).

To evaluate the specific capacitance (Cs) of the prepared ACs and nanocomposites, CV curves were recorded at different scan rates in three different electrolytes such as 1.0 M KOH, 1.0 M H2SO4, and 1.0 M Na2SO4; the results are presented in Fig. 8 and 9. As can be seen in Fig. 7, the shape of the CV curves is slightly distorted from the rectangular shape. The polarization resistance of the samples is the main reason for the distortion of CV curves.32 In addition, the presence of functional groups contributes to the pseudocapacitance behavior and the appearance of wide redox peak.33 Fig. 8 shows the calculated Cs values of ACs and nanocomposites at scan rate of 5 mV s−1. The superior capacitance performance of MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) was realized from the higher Cs values in all the three different electrolytes. At the scan rate of 5 mV s−1, the MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) achieved a maximum Cs of 410 F g−1, 345 F g−1 and 291 F g−1 in 1.0 M H2SO4, 1.0 KOH and 1.0 Na2SO4, respectively. However, under same conditions, the pure ACs showed the Cs values of 275 F g−1, 171 F g−1 and 22 F g−1 in 1.0 M H2SO4, 1.0 KOH and 1.0 Na2SO4, respectively. In comparison to pure ACs, the better Cs of MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) is due to the presence of MnO2. In fact, the charge can store on both MnO2 and ACs support, and therefore, the better Cs has been achieved.34 To the best of our knowledge, this is the first MnO2-based carbon nanocomposite to show good Cs in three kinds of electrolytes. Effect of MnO2 loading on electrochemical performance was also investigated. Fig. 8 represents the Cs values of nanocomposites with different MnO2 loading. As can be seen in Fig. 8, the loading of MnO2 has a profound effect on the electrochemical performance of the MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12). At higher loading of MnO2 [such as MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2)], the MnO2 has played a negative role by blocking the carbon surface and hinder the transport of ions.35 In addition, the blocking of interconnected pores of the carbon support by the MnO2 leads to a lower surface area of the nanocomposites and, therefore, the lower Cs values are obtained. Alike, very low MnO2 loadings also exhibited low Cs values which may be due to the insufficient amount of MnO2.36 This result agrees well with the BET results.


image file: c5ra16624a-f7.tif
Fig. 7 Cyclic voltammetry measurements of pure ACs (a) and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) (b) in (i) 1.0 M H2SO4, (ii) 1.0 M KOH and (iii) 1.0 M Na2SO4 over a potential range from −1.0 to 1.0 V at a scan rate of 5 mV s−1.

image file: c5ra16624a-f8.tif
Fig. 8 Specific capacitances of (a) pure ACs, (b) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2), (c) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]1), (d) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]2), (e) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]4), (f) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]8), (g) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]10), (h) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12), (i) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]14), and (j) MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]16) at scan rate of 5 mV s−1.

Further, the Cs of pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) at different scan rates was measured in 1.0 M H2SO4; the results are presented in Fig. 9. It was noticed that the Cs values are gradually decreased with scan rate for both ACs and MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y). At the scan rate of 500 mV s−1, the ACs showed nearly zero capacitance, whereas, the MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) maintained a maximum Cs of about 50 F g−1. The EIS spectrum [Fig. 10(i)] of ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) presented a depressed semicircle and a smaller interfacial charge-transfer resistance, representing good conductivity of the materials and high ion transfer speed across interfaces between the electrolyte and electrode.37 Fig. 10(ii) represents the cycle stability of pure ACs and MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) at scan rate of 5 mV s−1 in 1.0 M H2SO4.


image file: c5ra16624a-f9.tif
Fig. 9 Specific capacitance of (a) pure ACs and MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) at different scan rates determined in 1.0 H2SO4.

image file: c5ra16624a-f10.tif
Fig. 10 (i) Nyquist plots of (a) pure ACs and MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) recorded in 1.0 M H2SO4 and (ii) cycle stability of pure ACs (a) and MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) (b) in 1.0 M H2SO4; inset: galvanostatic charge–discharge cycles.

It was calculated that about 70% of the Cs was maintained by the MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) after 500 cycles. The result confirms the good capacitance retention of the electrode material. However, after 500 cycles, the Cs of pure ACs was decreased to 59%. Overall, the better performance of the MnO2/ACs(X[thin space (1/6-em)]:[thin space (1/6-em)]Y) is due five obvious reasons (i) the high specific surface area, (ii) high average pore size and mean pore volume, (iii) the presence of heteroatom such as O, S and N, (iv) conductivity of carbon and (v) needle-like MnO2.

The Cs performance of the present MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) is better or comparable to many other MnO2-based electrode materials (see Table 2) including graphene–MnO2, MnO2/activated CNTs, C/MnO2 nanorods, MnO2/ACs, MnO2/aniline, MnO2–VACNF, MnO2/diamond, flexible carbon cloth based MnO2 nanosheets, ternary MnO2/graphene nanosheets/CNTs composites, and MnO2 nanocrystals/carbon spheres.

Table 2 Comparison of the capacity for different MnO2-based electrode materials
Nanocomposites Cs F g−1 Scan rate mV s−1 Reference
MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) 410 (1.0 H2SO4) 5 This work
345 (1.0 KOH)
291 (1.0 Na2SO4)
Graphene–MnO2 310 2 16
MnO2/activated CNTs 250 10 38
C/MnO2 nanorods 165 5 17
MnO2/ACs 62 39
MnO2/aniline 626 10 40
MnO2–VACNF 437 1 41
MnO2/diamond 326 10 42
Flexible carbon cloth based MnO2 nanosheets 683.73 43
Ternary MnO2/graphene nanosheets/CNTs composite 367 20 44
MnO2/carbon spheres 412 2 45


4. Conclusions

In summary, needle-like MnO2 with different weight percentage was successfully decorated on ACs by a simple wet impregnation method. The physicochemical properties of the prepared nanocomposites were characterized in detail. For the first time, we demonstrated the MnO2-based carbon nanocomposite as electrode materials for supercapacitor in three different electrolytes (1.0 M H2SO4, 1.0 M KOH and 1.0 M Na2SO4). The nanocomposites were worked well as electrode materials for supercapacitors. Interestingly, the MnO2/ACs achieved a maximum capacitance of 410 F g−1, 345 F g−1 and 291 F g−1 in 1.0 M H2SO4, 1.0 KOH and 1.0 Na2SO4, respectively. Moreover, the MnO2/ACs(1[thin space (1/6-em)]:[thin space (1/6-em)]12) showed capacitance of ∼300 F g−1 even after 500 cycles in 1.0 M H2SO4.

Notes and references

  1. P. Simon and Y. Gogotsi, Nat. Mater., 2008, 7, 845 CrossRef CAS PubMed.
  2. D. N. Futaba, K. Hata, T. Yamada, T. Hiraoka, Y. Hayamizu, Y. Kakudate, O. Tanaike, H. Hatori, M. Yumura and S. Iijima, Nat. Mater., 2006, 5, 987 CrossRef CAS PubMed.
  3. L. F. Chen, X. D. Zhang, H. W. Liang, M. Kong, Q. F. Guan, P. Chen, Z. Y. Wu and S. H. Yu, ACS Nano, 2012, 6, 7092 CrossRef CAS PubMed.
  4. H. P. Cong, X. C. Ren and P. Wang, Chem. Soc. Rev., 2009, 38, 2520 RSC.
  5. W. Qian, F. Sun, Y. Xu, L. Qiu, C. Liu, S. Wang and F. Yan, Energy Environ. Sci., 2014, 7, 379 CAS.
  6. G. Ma, D. Guo, Q. Yang, X. Zhou, X. Zhao and Z. Q. Lei, RSC Adv., 2015, 5, 64704 RSC.
  7. M. Biswal, A. Banerjee, M. Deo and S. Ogale, Energy Environ. Sci., 2013, 6, 1249 CAS.
  8. C. Peng, X. Yan, R. Wang, J. Lang, Y. Ou and Q. Xue, Electrochim. Acta, 2013, 87, 401 CrossRef CAS PubMed.
  9. J. Zhang, L. Gong, K. Sun, J. Jiang and X. Zhang, J. Solid State Chem., 2012, 16, 2179 CAS.
  10. M. P. Bichat, E. Raymundo-Piñero and F. Beguin, Carbon, 2010, 48, 4351 CrossRef CAS PubMed.
  11. H. Li, R. Wang and R. Cao, Microporous Mesoporous Mater., 2008, 111, 32 CrossRef CAS PubMed.
  12. K. Lota, A. Sierczynska and G. Lota, Int. J. Electrochem., 2011, 2011, 321473 Search PubMed.
  13. J. Jiang and A. Kucernak, Electrochim. Acta, 2002, 47, 2381 CrossRef CAS.
  14. M. Toupin, T. Brousse and D. Belanger, Chem. Mater., 2004, 16, 3184 CrossRef CAS.
  15. G. Yu, L. Hu, N. Liu, H. Wang, M. Vosgueritchian, Y. Yang, Y. Cui and Z. Bao, Nano Lett., 2011, 11, 4438 CrossRef CAS PubMed.
  16. J. Yan, Z. Fan, T. Wei, W. Qian, M. Zhang and F. Wei, Carbon, 2010, 48, 3825–3833 CrossRef CAS PubMed.
  17. R. K. Sharma, H. S. Oh, Y. G. Shul and H. Kim, J. Power Sources, 2007, 173, 1024 CrossRef CAS PubMed.
  18. M. Zhi, C. Xiang, J. Li, M. Li and N. Wu, Nanoscale, 2013, 5, 72–88 RSC.
  19. J. P. Paraknowitsch and A. Thomas, Energy Environ. Sci., 2013, 6, 2839 CAS.
  20. K. Lee, D. Kim, Y. Yoon, J. Yang, H. G. Yun, I. K. You and H. Lee, RSC Adv., 2015, 5, 60914 RSC.
  21. Y. Zhang, J. P. Small, W. V. Pontius and P. Kim, Appl. Phys. Lett., 2005, 86, 073104 CrossRef PubMed.
  22. M. Wu, G. A. Snook, G. Z. Chen and D. J. Fray, Electrochem. Commun., 2004, 6, 499 CrossRef CAS PubMed.
  23. M. Gopiraman, S. G. Babu, Z. Khatri, W. Kai, Y. A. Kim, M. Endo, R. Karvembu and I. S. Kim, J. Phys. Chem. C, 2013, 117, 23582 CAS.
  24. M. Gopiraman, S. G. Babu, Z. Khatri, W. Kai, Y. A. Kim, M. Endo, R. Karvembu and I. S. Kim, Carbon, 2013, 62, 135 CrossRef CAS PubMed.
  25. X. Li, G. Zhang, X. Bai, X. Sun, X. Wang, E. Wang and H. Dai, Nat. Nanotechnol., 2008, 3, 538 CrossRef CAS PubMed.
  26. A. Cuesta, P. Dhamelincourt, J. Laureyns, A. Martínez-Alonso and J. M. D. Tascon, Carbon, 1994, 32, 1523 CrossRef CAS.
  27. M. Gopiraman, R. Karvembu and I. S. Kim, ACS Catal., 2014, 4, 2118 CrossRef CAS.
  28. L. Sun, C. Tian, Y. Fu, Y. Yang, J. Yin, L. Wang and H. Fu, Chem.–Eur. J., 2014, 20, 564 CrossRef CAS PubMed.
  29. T. Tojo, K. Sakurai, H. Muramatsu, T. Hayashi, K. S. Yang, Y. Jung, C. M. Yang, M. Endo and Y. Kim, RSC Adv., 2014, 4, 62678 RSC.
  30. T. C. Nagaiah, S. Kundu, M. Bron, M. Muhler and W. Schuhmann, Electrochem. Commun., 2010, 12, 338 CrossRef CAS PubMed.
  31. X. Lu, T. Zhai, X. Zhang, Y. Shen, L. Yuan, B. Hu, L. Gong, J. Chen, Y. Gao, J. Zhou, Y. Tong and Z. Wang, Adv. Mater., 2012, 24, 938 CrossRef CAS PubMed.
  32. R. Jiang, T. Huang, Y. Tang, J. Liu, L. Xue, J. Zhuang and A. Yu, Electrochim. Acta, 2009, 54, 7173 CrossRef CAS PubMed.
  33. X. Fan, Y. Lu, H. Xu, X. Kong and J. Wang, J. Mater. Chem., 2011, 21, 18753 RSC.
  34. S. Ghasemi, R. Hosseinzadeh and M. Jafari, Int. J. Hydrogen Energy, 2015, 40, 1037 CrossRef CAS PubMed.
  35. Y. He, W. Chen, X. Li, Z. Zhang, J. Fu, C. Zhao and E. Xie, ACS Nano, 2013, 7, 174 CrossRef CAS PubMed.
  36. Y. Peng, Z. Chen, J. Wen, Q. Xiao, D. Weng, S. He, H. Geng and Y. Lu, Nano Res., 2011, 4, 216 CrossRef CAS.
  37. D. Antiohos, K. Pingmuang, M. S. Romano, S. Beirne, T. Romeo, P. Aitchisonc, A. Minett, G. Wallacea, S. Phanichphant and J. Chen, Electrochim. Acta, 2013, 101, 99 CrossRef CAS PubMed.
  38. J. M. Ko and K. M. Kim, Mater. Chem. Phys., 2009, 114, 837 CrossRef CAS PubMed.
  39. A. Yuan and Q. Zhang, Electrochem. Commun., 2006, 8, 1173 CrossRef CAS PubMed.
  40. R. I. Jafri, A. K. Mishra and S. Ramaprabhu, J. Mater. Chem., 2011, 21, 17601 RSC.
  41. S. A. Klankowski, G. P. Pandey, G. Malek, C. R. Thomas, S. L. Bernasek, J. Wub and J. Li, Nanoscale, 2015, 7, 8485 RSC.
  42. S. Yu, N. Yang, H. Zhuang, J. Meyer, S. Mandal, O. A. Williams, I. Lilge, H. Schoonherr and X. Jiang, J. Phys. Chem. C, 2015, 119, 18918 CAS.
  43. S. He and W. Chen, J. Power Sources, 2015, 294, 150 CrossRef CAS PubMed.
  44. M. Ramezani, M. Fathi and F. Mahboubi, Electrochim. Acta, 2015, 174, 345 CrossRef CAS PubMed.
  45. X. Wang, X. Fan, G. Li, M. Li, X. Xiao, A. Yu and Z. Chen, Carbon, 2015, 93, 258 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available: SEM images, EDS spectra, BET isotherm curves and XRD spectra are provided. See DOI: 10.1039/c5ra16624a

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