Next Article in Journal
Use of Carbon Nanoparticles to Improve Soil Fertility, Crop Growth and Nutrient Uptake by Corn (Zea mays L.)
Previous Article in Journal
Novel, Simple and Low-Cost Preparation of Ba-Modified TiO2 Nanotubes for Diclofenac Degradation under UV/Vis Radiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nest-Like MnO2 Nanowire/Hierarchical Porous Carbon Composite for High-Performance Supercapacitor from Oily Sludge

1
College of Mechanical and Electronic Engineering, Shandong University of Science and Technology, Qingdao 266590, China
2
College of New Energy, China University of Petroleum (East China), Qingdao 266580, China
3
State Grid Shandong Electric Power Research Institute, Jinan 250003, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(10), 2715; https://doi.org/10.3390/nano11102715
Submission received: 4 September 2021 / Revised: 10 October 2021 / Accepted: 12 October 2021 / Published: 14 October 2021

Abstract

:
In the aim to go beyond the performance tradeoffs of classic electric double-layer capacitance and pseudo-capacitance, composites made out of carbon and pseudo-capacitive materials have been a hot-spot strategy. In this paper, a nest-like MnO2 nanowire/hierarchical porous carbon (HPC) composite (MPC) was successfully fabricated by a controllable in situ chemical co-precipitation method from oily sludge waste. Due to the advantages of high surface area and fast charge transfer for HPC as well as the large pseudo-capacitance for MnO2 nanowires, the as-prepared MPC has good capacitance performance with a specific capacitance of 437.9 F g−1 at 0.5 A g−1, favorable rate capability of 79.2% retention at 20 A g−1, and long-term cycle stability of 78.5% retention after 5000 cycles at 5 A g−1. Meanwhile, an asymmetric supercapacitor (ASC) was assembled using MPC as the cathode while HPC was the anode, which exhibits a superior energy density of 58.67 W h kg−1 at the corresponding power density of 498.8 W kg−1. These extraordinary electrochemical properties highlight the prospect of our waste-derived composites electrode material to replace conventional electrode materials for a high-performance supercapacitor.

Graphical Abstract

1. Introduction

Electrical energy storage systems have drawn great attention in recent decades such as conventional capacitors and batteries, Li-ion batteries, and supercapacitors [1]. Among them, supercapacitors, also referred to as electrochemical capacitors, can be seen as the primary choice for applications in energy storage systems because of their higher energy density than conventional capacitors and better power density than batteries [2]. Generally, supercapacitors can be classified into electric double-layer capacitors (EDLC) and pseudo capacitors, according to the different energy storage mechanisms. The capacitance in EDLCs is controlled by the surface electrical double layers between the electrode and electrolyte, while in pseudo capacitors, the energy is stored by the fast and reversible faradaic redox reactions in the electrode, which are controlled by diffusion. [3]. As is known to all, pseudo capacitors usually provide higher specific capacitances (Cs) and energy densities, but suffer from some shortcomings such as poor electrical conductivity and cycling stability, while EDLCs are the opposite. The plans to realize the synergistic effect of the advantages of both EDLC and pseudo capacitor are highly desirable to improve the electrochemical performance [4].
The electrode material plays a key role in the development of high-performance supercapacitors [5]. In recently reported literatures, various candidates have been investigated including carbon materials, conducting polymers, transition metal oxides, and hydroxides [6]. Among them, transition metal oxides have attracted considerable interest because of the variety of oxidation states for efficient redox charge transfer, which can provide pseudo-capacitance effects such as MnxOy [7,8], CoxOy [9], FexOy [10], and NixOy [11]. MnO2 is one of the most promising metal oxides due to its low cost, high theoretical capacitance, and environmental safety properties [12]. The capacitance can be significantly enhanced through fast faradaic redox reactions occurring on the surface of MnO2 [13]. However, as with other pseudo-capacitor electrode materials, MnO2 suffers from low surface area in bulk and poor electrical conductivity, which limits its practical capacitance application. Hence, an emerging new concept is to integrate MnO2 nanostructures with the conductive substrates to form a composite. Moreover, the interface interactions between the MnO2 nanostructure and substrates were of great influence on the capacitance performance of the composites [14].
With excellent electrical conductivity and high specific surface area, carbon nanomaterials such as graphene [15], carbon nanofibers [16], carbon nanotubes [17], and their hybrids [18,19,20] have been selected as substrate components for preparing MnO2-carbon composites, which not only provide stable scaffolds for MnO2, but also facilitate charge transport for pseudo capacitive reactions of MnO2 during the charge/discharge process [21,22]. However, these nano-sized composites are still limited in practical application by problems associated with their high cost. Meanwhile, given the strong van der Waals interactions, the nano-sized materials are easy to stack together during the electrode preparation [23], which may lead to sacrificing the effective surface area and the reduction in specific capacitance [24]. Recent studies have shown that 3D hierarchical nanostructured carbons are promising candidates as substrates of nanoscaled MnO2 accumulation to fabricate hybrid composites [25,26] because of their unique internal porous structures. For example, 3D graphene foam [27], carbon nanotube sponges [28], hierarchical porous carbon (HPC) [29,30], and hierarchical hybrid nanostructures [31,32] have been developed. Despite these achievements, it is also desirable to develop simple and scalable methods to synthesize MnO2/HPC composites. HPCs are usually prepared from polymer precursors using template-assisted methods. In recent years, a lot of waste materials and biomass resources such as plant fibers [30,33], animal tissues [34], or industry waste [35] have been transformed into HPCs by their self-structuring procedure. Obviously, the high value-added utilization of worthless and renewable material is beneficial to both environmental protection and sustainable development [36,37]. As the most refractory contaminants in the petroleum industry, oily sludge (OS) has become a great threat to the environment [38]. To our knowledge, few works have been carried out so far concerning about OS-based HPCs and MnO2 composites.
In this work, an effective approach was implemented to synthesize the MnO2/HPC composite material (MPC) by using OS waste. First, the HPC substrate was prepared by carbonization and the chemical activation method. Then, the obtained HPC was surface modified by an acid pre-oxidation process. Furthermore, MnO2 nanowires were grown on the HPC substrate using a controllable in situ chemical co-precipitation method. The MnO2/HPC nanostructured electrode has been used as an anode in supercapacitors and has demonstrated high capacitance and excellent cycling stability. HPC served as the mechanical support to avoid structure collapse and provided highways for electronic and ionic transport to MnO2 nanowires. MnO2 nanowires enhanced the specific capacitance by introducing reversible redox reactions at the same time.

2. Materials and Methods

2.1. Materials

The oily sludge sample used in this study was a conventional oil tank bottom sludge in the petrochemical industry. It was obtained during the tank cleaning process in the Dongying petroleum storage depot of SINOPEC. The OS was filtered and dried at 105 °C overnight as the precursor of this study. The OS precursor contains 32.2% of oil organic components, 16.6% of moisture, and the rest is solid residue. All the other chemical reagents used in this work were of AR grade and used as received without further purification.

2.2. Preparation of MPCs

The preparation process of MPC is illustrated in Figure 1. First, OS-derived HPC material was prepared by carbonization and the chemical activation method according to our previous work [39]. Briefly, oily sludge was carbonized at 700 °C under N2 flow. After being washed by HF to remove inorganic impurities, the carbon residue was activated by KOH at a mass ratio of 1:3 under 700 °C for 2 h. Second, the as-received HPC was pre-oxidized in 2 M HNO3 for 6 h to obtain HPC-A. Then, 200 mg HPC-A and 790 mg KMnO4 powder were ultrasonically mixed in 100 mL of pure water in a flask. The flask was heated and refluxed at different times at 70 °C in a water bath under intense stirring. The precipitates were filtered out and rinsed with pure water for several times until the filtrate became colorless. The MPC samples synthesized under different refluxing times of 1 h, 3 h, and 5 h were named MPC-A-1, MPC-A-3, and MPC-A-5, respectively. As for a control sample, MPC-3 sample was synthesized following the same procedure, except that HPC was used instead of HPC-A.

2.3. Characterization

The morphology of as-prepared materials was analyzed by scanning electron microscopy (SEM, JSM-6700F, JEOL, Tokyo, Japan). Transmission electron microscopy (TEM) and energy dispersive spectrum (EDS) mapping images of samples were acquired on a JEM-2100UHR (JEOL, Tokyo, Japan) at an accelerating voltage of 100 kV. X-ray diffraction (XRD) patterns were recorded by PANalytical X’Pert Pro (Panalytical, Almelo, Netherland) with Cu Kα radiation (λ = 1.54178 Å). The surface chemical states of samples were characterized by X-ray photoelectron spectra (XPS) on an ESCALAB MK II (Thermo Scientific, Waltham, MA, USA) with Mg Kα (hυ = 1253.6 eV) as the excitation source. The pore structures were tested by a Multipoint N2 adsorption–desorption experiment on an automatic Micromeritics ASAP 2020 (Micromeritics, Norcross, GA, USA) analyzer at 77 K. The specific surface area was calculated by the BET method and the pore size distribution was generated from the desorption branch of the isotherm by the non-local density functional theory (NLDFT) method.

2.4. Electrochemical Measurements

To fabricate the working electrode, as-prepared MPCs samples were first ground below 45 μm and then mixed with carbon black and polyvinylidene-fluoride (PVDF) with a ratio of 8:1:1 (active material:conductive agent:binder) in N-methyl-2-pyrrolidone (NMP) to form a homogeneous slurry. Then, the resulting slurry was coated on nickel foam as a 1 cm × 1 cm sheet, followed by drying at 80 °C overnight. The mass loading of active materials on electrodes was controlled at about 2.8, 3.0, 2.8, 2.6, and 4.2 mg cm−2 for the MPC-A-1, MPC-A-3, MPC-A-5, MPC-3, and HPC-A electrodes, respectively.
Cyclic voltammetry (CV) and galvanostatic charge/discharge (GCD) electrochemical tests were carried out on a CHI660E electrochemical workstation. Electrochemical impedance spectroscopy (EIS) were performed on the Gramy Reference 600 (Shanghai Chenhua Science Technology Corp., Ltd., Shanghai, China) with a frequency range of 105 to 10−2 Hz at an open circuit voltage. For typical three-electrode system measurements, an Ag/AgCl electrode and platinum sheet were employed as the reference and counter electrodes, respectively. Potential range for the CV and GCD test was from 0 to 1 V. For two-electrode system measurements, the MPC and HPC-A electrode was used as the cathode and anode, respectively, and then separated by a thin polypropylene film to fabricate the ASC device. The electrolyte was a 1 M Na2SO4 solution for both three-electrode and two-electrode systems.
The charge balance and the mass ratio of the cathode and anode in the ASC device and the detailed calculation of the specific capacitance of the electrodes and the energy density and power density of the as-assembled ASC device are shown in the Supplementary Materials.

3. Results and Discussion

3.1. Characterizations of MPC Samples

As shown in Figure S1a,b, a cross-linked porous structure could be found in the carbon skeleton in HPC-A, which maintains its hierarchical porous structure of HPC (Figure S1c,d) [39] after oxidation treatment by HNO3. The SEM images of MPCs are given in Figure 2 to analyze the structure details of the materials. The porous structure in the HPC-A substrate that provides space for MnO4 ions to enter the interior and be reduced to MnO2 on the interior and exterior surfaces follows the equation [40]:
4   MnO 4 + 3   C + H 2 O     4   MnO 2 + CO 3 2 + 2   HCO 3
In addition, the content and morphology of MnO2 in the as-received MPC could be controlled by varying the reaction time. As shown in Figure 2a, a unique “nest-like” porous structure with compact MnO2 nanowires with anisotropic growth could be observed on the surface of the carbon matrix in MPC-A-1. With an increase in reaction time, the diameters of MnO2 nanowires were increased, as shown in the SEM images of MPC-A-3 (Figure 2b), which made the “nest” more tight. As for MPC-A-5 (Figure 2c), the hierarchical porous structure of the HPC-A substrate was almost completely covered by the surface MnO2 nanostructure due to the over-reaction between carbon and KMnO4. Moreover, some egg-like structure could be observed in all of these samples, which may have been caused by the shrinking of the defective carbon materials when reacted with the strong oxidant KMnO4. Compared with MPC-A-3, the MPC-3 samples that were obtained under the same reaction time possessed shorter and more disordered arranged MnO2 nanowires that nearly could not form a “nest” (Figure 2d). This shows that the HPC-A substrate could make MnO2 nanowire anchoring more uniform, and effectively promote the growth of MnO2 nanowires during the following redox reaction. The MPC-A-3 possesses the structure of stable MnO2 nanowires wrapped on a HPC substrate, which can promote the pseudo capacitive of MnO2 as well as EDLC effect of porous carbon to increase the capacitance performance.
TEM images of MPC-A-3 also showed a nest-like structure, where nanowires with a diameter below 10 nm were covered on the surface of the HPC-A substrate in Figure 3a,b. Furthermore, the results of the corresponding X-ray elemental mapping images of C, Mn, and O (Figure 3c–f) demonstrate the co-existence of these elements. Combined with the above SEM images, the MnO2 nanowires were successfully embedded in the HPC structures by the above method.
The composition of the samples was examined through the XRD test. As shown in Figure S1e, the HPC-A sample had a disordered and defective graphitic structure as evidently illustrated by the appearance of two broad peaks of (002) and (101) diffraction planes. The XRD patterns of MPCs are shown in Figure 4a, in which the broad diffraction peaks around 26° and 44° indicate the presence of amorphous graphitic carbon in the composites [41], while the characteristic peaks at 12°, 37°, and 66° correspond to the (001), (111), and (020) crystal planes of birnessite-type MnO2 (JCPDS 42-1317) [42], respectively. It was noticed that the intensity of the diffraction peak at 37° gradually increased with the reaction time due to the greater amount of MnO2 in the composites.
TGA tests were performed in an air atmosphere from 25 to 600 °C to calculate the loading amount of MnO2 in different MPCs. As shown in Figure 4b, the total weight loss of the HPC-A material was 95.45% and the remaining residue was mainly the unburned inorganic impurities. In addition, the weight loss of MPC-A-1, MPC-A-3, and MPC-A-5 was 70.24%, 61.72%, and 57.61%, respectively, which decreased with the increase in the reaction time. It is clear that the weight losses of the MPCs are mainly due to the combustion of carbon components. The MnO2 loading amount can be calculated by the difference of weight loss between HPC-A and MPCs. Therefore, the mass percentage of MnO2 contained in the MPC-A-1, MPC-A-3, and MPC-A-5 composites can be calculated as 26.41%, 35.34%, and 39.64%, respectively, which indicates that a longer reaction time was beneficial to MnO2 loading.
Combined with the observations in the SEM and TEM images, different MnO2 loading amounts could affect the porosity of MPCs. The N2 adsorption–desorption results of MPCs and HPCs are illustrated in Figure 4c and Figure S1f, respectively. All the samples exhibited a typical combined I/IV type isothermal with an obvious H4 hysteresis loop [43]. As is well-known, the sharp increases in adsorption capacity at low relative pressure are related to the micropores, while the relatively slower increases are caused by the capillary condensation of N2 in the mesopores of the sample. This means that the surface MnO2 loading did not change the basic skeleton structure of the total materials.
The pore size distributions of the MPCs are shown in Figure S2, and the pore structure parameters are listed in Table S1. The specific surface area (SSA) of the MPC sample is closely related to the reaction time. With the increase in reaction time from 1 h to 3 h and 5 h, the SSA decreased from 1957.4 to 1437.1 and 938.4 m2 g−1, respectively, which mainly resulted from the pore blocking. The total pore volume decreased from 1.397 to 0.775 cm3 g−1, which means that the MnO2 nanostructure is generated and filled in the pores of HPC-A. The increase in loading amount would cause more blockage of pores.
To further investigate the effect of pre-oxidation on MnO2 loading, FTIR analyses were performed, as shown in Figure 4d. The peaks of around 1100 cm−1, 1580 cm−1, and 3440 cm−1, which represent –C–C/–C–O, –C=C, and –OH, can clearly be seen in the spectra of the origin HPC sample, respectively. After being oxidized by HNO3, it was found that the characteristic peak derived from symmetric carboxyl appeared at 1680 cm−1, which indicates the acidic oxygen-containing surface functional groups’ oxidation treatment [44]. The FTIR spectra for MPC-A-3 shows the vibration of the Mn–O bond at the wavelength of 547 cm−1 in the fingerprint region. Moreover, it is known that the positions of symmetric –COO stretching bonds considerably vary in the range of 1300–1450 cm−1 in metal complexes depending on the coordination structures [45]. The appearance of a new peak at 1380 cm−1 is indicative of the symmetric carboxylate stretching mode, which also represents the existence of MnO2. Furthermore, the peak at 1680 cm−1 assigned to –C=O in HPC-A showed a red shift in MPC-A-3. This indicates that the double bond between carbon and oxygen on the HPC-A substrate becomes longer and weaker due to the formation of the electrostatic ionic bond with MnO2.
The preparation mechanism of MPC materials was analyzed according to the above characterization results. As shown in Figure 5, abundant carboxyl and hydroxyl groups were produced on the surface of HPC-A. On one hand, these oxygen-containing functional groups made the reaction solution acidic, which could enhance the oxidation effect of KMnO4 and increase the reaction depth of KMnO4 and carbon [46]. On the other hand, these groups were beneficial in the generation of stronger chemical bond interactions between newly formed MnO2 and the surface of the carbon substrates. The nanowires were formed via head-to-head overlapping of nucleated MnO2 nanoparticles [47]. The head-to-head overlapping is possible because the bonding sites on the HPC surface are quite small, and the high number of coordination sites trigger the preferential condensation of the nucleated MnO2 nanoparticles into long ultrathin nanowires. Then, the slow growth and reduction in the MnO2 nanowires proceeded simultaneously during the reaction. Finally, nanowires led to further growth, aggregation, and assembled into a 3D nest-like structure due to the narrow space in the porous carbon substrates. Last but not least, the large SSA of the HPC-A substrate could provide a large number of sites for the embedding of MnO2 nanowires.

3.2. Capacitance Performance of MPCs

The electrochemical performance of all MPCs samples was first investigated in a three-electrode system in a Na2SO4 aqueous solution at room temperature. The representative CV curves recorded at a scan rate of 20 mV s−1 and a potential window from 0 to 1 V are shown in Figure 6a. All curves displayed an approximate rectangular shape, demonstrating the typical EDLC effect of electrode materials. Although redox peaks could not be clearly found in the CV curves, this electrochemical signature is consistent with the pseudo-capacitance nature of MnO2, which arises from reactions that are faradaic in origin, involving the passage of charge across the double layer [13]. This pseudo-capacitance is based on the insertion of cations (H+ or Na+) from the electrolyte with MnO2 upon reduction followed by de-insertion upon oxidation:
MnO2 + H+ + e → MnOOH
MnO2 + Na+ + e → MnOONa
Among these curves, the MPC-A-3 electrode revealed the largest area of the CV triangle, suggesting its highest capacitance value, which was due to the large surface area, optimal porous structure, and MnO2 loading. Compared with the MPC-3 sample, a higher capacitance value of MPC-A-3 means that the pre-oxidation treatment of the carbon substrate is beneficial to the capacitance performance of the composites. This is because the HPC-A substrate provides more sites for MnO2 loading and regulates the growth of MnO2 nanowires in the axial direction. The uniform distribution of MnO2 on the HPC skeleton promotes the transfer of electrolyte ions, which improves the capacitance performance. In addition, the capacitance of MPC-A-x first increased with reaction time and then decreased because the overloaded MnO2 would lead to the blockage and destruction of the pore structure of composites, which inhibits the acceptance of capacitance active sites. The CV curves of MPC-A-3 at different scan rates in Figure 6b still exhibited a symmetrical shape even at a high scan rate of 100 mV s−1, demonstrating a faster and reversible charging and discharging process. Similar trends can be found in the CV curves of MPC-A-1, MPC-A-5, and MPC-3 at different scan rates (Figure S3). The outstanding electrochemical capacitance performance of MPCs can be attributed to the hierarchical porous structure with large SSA and unblocked ion transport path. The reproducibility performance of the results were tested by a typical CV test. As shown in Figure S4, the shape of the CV curves showed no significant changes compared with the original tests with the differences under 5%.
The Cs of MPCs was investigated by GCD tests. The GCD curves of MPC-A-3 presented in Figure 6c and that of the other MPCs (Figure S5) exhibited similar triangular and symmetrical forms. The Cs of the MPC electrodes was calculated from these GCD curves and is shown in Figure 6d. The Cs of MPCs followed the order of MPC-A-3 > MPC-3 > MPC-A-1 > MPC-A-5 at each current density. The Cs of MPC-A-3 could reach 437.9 F g−1 (1.31 F cm−2) at the current density of 0.5 A g−1 (1.5 mA cm−2). In general, with the increase in current densities, the capacitance will decrease along with a lower amount of electrolyte ions occupying the active sites. The Cs of MPC-A-3 could maintain 346.6 F g−1 (1.04 F cm−2) at 20 A g−1 (60 mA cm−2) with the rate capability of 79.2%, which is much higher than MPC-A-1 (74.4%), MPC-A-5 (40.5%), and MPC-3 (73.8%). It is worth noting that MPC-3 showed a slightly worse rate capability than MPC-A-3 despite the same synthesis time. This could be due to a small number of MnO2 nanowires possibly falling off from the surface of the MPC-3 at the high charging/discharging current densities, which are caused by the weaker binding effort between MnO2 nanowires and the HPC substrate without pre-oxidized treatment. Furthermore, for MPC-A-3, the cycling stability performance was evaluated by the GCD test at the current density of 5 A g−1 for 5000 cycles. As illustrated in Figure 6e, the capacitance increased slightly in the first 500 cycles due to the initial activation process of MnO2 materials for pseudo capacitance. After 5000 cycles, the capacitance retention remained at 78.5% of the initial value, indicating outstanding cycling stability. The inset in Figure 6e shows the GCD curves after the 3000th cycle. The symmetrical shape of the curves and the voltage drop were not changed significantly, although the total charging/discharging time became shorter, which related to worse capacitance performance. The good cycling stability of MPC-A-3 makes it prospective for the practical applications of a supercapacitor.
The Nyquist plots of the MPCs electrodes obtained by the EIS test are shown in Figure 6f, where the almost vertical curves at the low frequency are closely related to ionic diffusion resistance, and the semicircular at high-middle frequency is associated with charge transfer resistance [48]. Based on the equivalent circuit as inset in Figure 6f, the impedance data were fitted. It was observed that the semicircle increased in the order of MPC-3 (7.32 Ω) > MPC-A-5 (6.68 Ω) > MPC-A-1 (5.28 Ω) > MPC-A-3 (4.36 Ω), indicating the MPC-A-3 possesses the smallest interfacial charge-transfer resistance (Rct), which represents the optimal ionic diffusion capacity. The equivalent serials resistance (Rs) of each electrode followed the order of MPC-A-5 (5.36 Ω) > MPC-3 (3.38 Ω) > MPC-A-3 (1.15 Ω) > MPC-A-1 (0.94 Ω). Here, Rs increased with the increase in the reaction time of preparation, which is probably because the increase in MnO2 loading can inhibit the conductivity of the capacitive active material. Additionally, MPC-A-3 showed a lower Rs than MPC-3, which is due to the fact that the MnO2 combination onto MPC-A-3 was more ordered. This verifies that the pre-oxidation process could effectively improve the growth of MnO2 nanowires on the HPC-A substrate.
To elucidate the mechanism of the MPC composite electrode, the relative contributions of the EDLC and the pseudo capacitance to the electrochemical capacitance were further evaluated via CV curves. The current response at various potentials can be calculated by the following equation [49]:
i = a   ×   v b
where i(v) is the current density (mA); ν is the scan rate (mV s−1); and a and b are the coefficients. As it is known that the electrode shows a battery type with fully faradaic effect if b equals 0.5, then a total EDLC property has a b value of 1. The b values between 0.5 and 1 indicate the combination of the double layer capacitance and diffusion-controlled pseudo capacitance. As shown in Figure 7a, the cathodic and anodic b values for MPC-A-3 at a potential of 0.5 V are 0.8543 and 0.8504, respectively, implying the current response results from both surface EDLC capacitive and diffusion-controlled pseudo capacitive [50].
Moreover, the surface EDLC contribution and the pseudo capacitive contribution for MPC-A-3 electrode were also be calculated from the CV curves (Figure 6b) using the function [51]:
i =   k 1 v   + k 2 v 1 / 2
where k1 and k2 are the appropriate coefficient. The current response from the EDLC behavior could be determined by k 1 v . k1 equals the slope of each curve at different potentials like that illustrated in Figure 7b. Taking k 1 v as the ordinate and corresponding potential as the abscissa, the EDLC capacitance is obtained. As shown in Figure 7c, for instance, the capacitive contributions from different processes at a scan rate of 20 mV s−1 were determined. The surface EDLC capacitive were marked as the mesh area, and the rest of the total capacitance was contributed by the diffusion-controlled pseudo capacitive. The relative contributions at different scan rates are summarized in Figure 7d. The contribution of the EDLC process increases from 49.8% to 75.4% for scan rates from 10 to 100 mV s−1. At low scan rates, the dominant current response arises from the diffusion-controlled process, which attributed to the pseudo capacitance of MnO2. At a higher scan rate, the faradaic process is inhibited and the surface EDLC process plays the major role in charge storage.
To further evaluate the practical application of the MPC materials, ASC devices with extended working voltages was assembled with MPC-A-3 as the cathode and HPC-A as the anode. The electrochemical result of HPC-A is shown in Figure S6, which illustrates typical capacitance properties. According to the CV comparison in Figure S7, The optimal mass ratio of the cathode and anode was calculated as 0.68 [52]. As shown in the CV curves of this ASC device in Figure 8a, the potential window could be enlarged to 0–2.0 V. Nearly symmetric shape CV curves were obtained at different scan rates in the voltages of 2.0 V (Figure 8b). Afterward, GCD curves at different current densities with the voltage of 2.0 V are shown in Figure 8c. The curves exhibited a nearly linear potential-time behavior without significant voltage drop, suggesting rapid charging/discharging performance, which is consistent with the results of the CV tests. The specific capacitance can be calculated as 105.6, 91.5, 84.2, 79.8, 76.9, 70.8, and 60.2 F g−1 at 0.5, 1, 2, 3, 5, 10, and 20 A g−1 based on the total mass of two electrodes, indicating a good rate capability (Figure 8d). The area specific capacitance was 0.76, 0.66, 0.61, 0.57, 0.55, 0.51, and 0.43 F cm−2 at 3.6, 7.2, 14.4, 21.6, 36, 72, and 144 F cm−2, respectively. Moreover, the ASC device could undergo 5000 cycles with a capacitance retention of 83.9% and excellent coulombic efficiency (Figure 8e). The insets of Figure 8e illustrate the 5th and 5000th cycle GCD curves during the cycle stability test. The shape of the curves barely changed, although the total charging/discharging time was decayed due to the structure damage of the electrode after a long-term duration. Meanwhile, the energy densities and power densities of ASC were calculated and compared with the other reported data on the Ragone plot in Figure 8f. The maximum energy density of 58.67 W h kg−1 was obtained at the corresponding power density of 498.8 W kg−1, which still retained 33.44 W h kg−1 at 20 kW kg−1. These superior energy density properties are competitive with the reported MnO2-based supercapacitor devices listed in Table 1.

4. Conclusions

In summary, an oily sludge-based MnO2 nanowires/HPC composite electrode was successfully fabricated through an easy method. The MPC-A-3 electrode material delivered a high specific capacitance of 437.9 F g−1 at the current density of 0.5 A g−1, favorable rate capability of 79.2% retention at 20 A g−1, and long-term cycle stability. These good electrochemical properties could be attributed to the unique nanostructures of the MPC composite, where the HPC substrate improved conductivity and high surface area for EDLC behavior, and MnO2 nanowires provided enhanced pseudo-capacitance. Moreover, an ASC device assembled with MPC-A-3 as a cathode while the HPC as anode could be cycled reversibly at a prolonged voltage of 0–2.0 V. This ASC device maintained 83.9% of initial capacitance after 5000 cycles with an energy density of 58.67 W h kg−1 at a corresponding power density of 498.8 W kg−1. These results demonstrate that our MnO2 nanowire/HPC composite material from OS would be a promising electrode in practical supercapacitor applications.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11102715/s1, Figure S1: The SEM (a) and TEM images (b) of HPC; SEM (c) and TEM (d) images of HPC-A; the XRD patterns of HPC and HPC-A (e), and the N2 adsorption–desorption isotherm (f) of the HPC and HPC-A samples; Figure S2: Pore size distribution and cumulative pore volume curves of MPC-A-1 (a), MPC-A-3 (b), MPC-A-5 (c), and MPC-3 (d) calculated by the NLDFT method; Figure S3: CV curves of MPC-A-1 (a), MPC-A-5 (b), and MPC-3 (c) at different scan rates; Figure S4: Reproduced CV tests (a) of the MPC-A-3r at different scan rates, the comparison of the CV curves (b) of the reproduced tests and the original typical tests; Figure S5: GCD curves of MPC-A-1 (a), MPC-A-5 (b), and MPC-3 (c) at different charge/discharge current densities and the comparison of GCD curves (d) of different MPCs samples at a current density of 2 A g−1; Figure S6: The CV curves of the HPC-A anode at different scan rates (a) and the GCD curves of the HPC-A anode at different current densities (b); Figure S7: Comparative CV curves of MPC-A-3 and HPC-A electrodes; Table S1: Pore structure parameters of MPCs.

Author Contributions

Conceptualization, X.L. and Z.W.; Methodology, X.L.; Software, Z.G.; Validation, X.L., Z.W. and D.H.; Investigation, X.L.; Data curation, Z.G.; Writing—original draft preparation, D.H. and X.L.; writing—review and editing, Z.W.; Project administration, Z.W.; Funding acquisition, X.L. and Z.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of Shandong Province (grant numbers ZR2019BEE023 and ZR2020QE199) and the Opening Fund of State Key Laboratory of Heavy Oil Processing (grant number SKLOP202003001).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Zhang, L.L.; Zhao, X.S. Carbon-based materials as supercapacitor electrodes. Chem. Soc. Rev. 2009, 38, 2520–2531. [Google Scholar] [CrossRef] [PubMed]
  2. Li, Q.; Liu, J.; Zou, J.; Chunder, A.; Chen, Y.; Zhai, L. Synthesis and electrochemical performance of multi-walled carbon nanotube/polyaniline/MnO2 ternary coaxial nanostructures for supercapacitors. J. Power Sources 2011, 196, 565–572. [Google Scholar] [CrossRef]
  3. Wang, S.Z.; Gai, L.G.; Jiang, H.H.; Guo, Z.Z.; Bai, N.N.; Zhou, J.H. Reduced graphene oxide grafted by the polymer of polybromopyrroles for nanocomposites with superior performance for supercapacitors. J. Mater. Chem. A 2015, 3, 21257–21268. [Google Scholar] [CrossRef]
  4. Xia, T.; Wang, Q.; Wu, W.; Ao, C.; Zheng, Z.; Lu, C.; Chen, Z.; Zhang, W. Fabrication and characterization of MnO2-Coated carbon fabrics from silk for shape-editable supercapacitors. J. Alloys Compd. 2021, 854, 157289. [Google Scholar] [CrossRef]
  5. Cheng, Y.; Lu, S.; Zhang, H.; Varanasi, C.V.; Liu, J. Synergistic effects from graphene and carbon nanotubes enable flexible and robust electrodes for high-performance supercapacitors. Nano Lett. 2012, 12, 4206–4211. [Google Scholar] [CrossRef]
  6. Wang, G.; Zhang, L.; Zhang, J. A review of electrode materials for electrochemical supercapacitors. Chem. Soc. Rev. 2012, 41, 797–828. [Google Scholar] [CrossRef] [Green Version]
  7. Dahal, B.; Mukhiya, T.; Ojha, G.P.; Chhetri, K.; Tiwari, A.P.; Muthurasu, A.; Lee, M.; Chae, S.-H.; Kim, T.; Chung, D.C.; et al. A multicore-shell architecture with a phase-selective (α+δ)MnO2 shell for an aqueous-KOH-based supercapacitor with high operating potential. Chem. Eng. J. 2020, 387, 124028. [Google Scholar] [CrossRef]
  8. Zhuang, R.; Dong, Y.; Li, D.; Liu, R.; Zhang, S.; Yu, Y.; Song, H.; Ma, J.; Liu, X.; Chen, X. Polyaniline-mediated coupling of Mn3O4 nanoparticles on activated carbon for high-performance asymmetric supercapacitors. J. Alloys Compd. 2021, 851, 156871. [Google Scholar] [CrossRef]
  9. Liu, X.B.; Wu, Z.P.; Yin, Y.H. Highly nitrogen-doped graphene anchored with Co3O4 nanoparticles as supercapacitor electrode with enhanced electrochemical performance. Synth. Met. 2017, 223, 145–152. [Google Scholar] [CrossRef]
  10. Song, Z.; Liu, W.; Wei, W.; Quan, C.; Sun, N.; Zhou, Q.; Liu, G.; Wen, X. Preparation and electrochemical properties of Fe2O3/reduced graphene oxide aerogel (Fe2O3/rGOA) composites for supercapacitors. J. Alloys Compd. 2016, 685, 355–363. [Google Scholar] [CrossRef]
  11. Liu, T.; Jiang, C.J.; Cheng, B.; You, W.; Yu, J.G. Hierarchical flower-like C/NiO composite hollow microspheres and its excellent supercapacitor performance. J. Power Sources 2017, 359, 371–378. [Google Scholar] [CrossRef]
  12. Wei, W.; Cui, X.; Chen, W.; Ivey, D.G. Manganese oxide-based materials as electrochemical supercapacitor electrodes. Chem. Soc. Rev. 2011, 40, 1697–1721. [Google Scholar] [CrossRef]
  13. Brousse, T.; Bélanger, D.; Long, J.W. To Be or Not To Be Pseudocapacitive? J. Electrochem. Soc. 2015, 162, A5185–A5189. [Google Scholar] [CrossRef] [Green Version]
  14. Li, Q.; Kuzmenko, V.; Haque, M.; Di, M.; Smith, A.D.; Lundgren, P.; Enoksson, P. Explanation of anomalous rate capability enhancement by manganese oxide incorporation in carbon nanofiber electrodes for electrochemical capacitors. Electrochim. Acta 2020, 340, 135921. [Google Scholar] [CrossRef]
  15. Cheng, Q.; Tang, J.; Ma, J.; Zhang, H.; Shinya, N.; Qin, L.-C. Graphene and nanostructured MnO2 composite electrodes for supercapacitors. Carbon 2011, 49, 2917–2925. [Google Scholar] [CrossRef]
  16. Zhang, J.; Zhao, X.; Huang, Z.; Xu, T.; Zhang, Q. High-performance all-solid-state flexible supercapacitors based on manganese dioxide/carbon fibers. Carbon 2016, 107, 844–851. [Google Scholar] [CrossRef]
  17. Lei, Z.; Shi, F.; Lu, L. Incorporation of MnO2-coated carbon nanotubes between graphene sheets as supercapacitor electrode. ACS Appl. Mater. Interfaces 2012, 4, 1058–1064. [Google Scholar] [CrossRef]
  18. Deng, L.; Hao, Z.; Wang, J.; Zhu, G.; Kang, L.; Liu, Z.-H.; Yang, Z.; Wang, Z. Preparation and capacitance of graphene/multiwall carbon nanotubes/MnO2 hybrid material for high-performance asymmetrical electrochemical capacitor. Electrochim. Acta 2013, 89, 191–198. [Google Scholar] [CrossRef]
  19. Bi, T.; Fang, H.; Jiang, J.; He, X.; Zhen, X.; Yang, H.; Wei, Z.; Jia, Z. Enhance supercapacitive performance of MnO2/3D carbon nanotubes-graphene as a binder-free electrode. J. Alloys Compd. 2019, 787, 759–766. [Google Scholar] [CrossRef]
  20. Liu, Y.; He, D.; Duan, J.; Wang, Y.; Li, S. Synthesis of MnO2/graphene/carbon nanotube nanostructured ternary composite for supercapacitor electrodes with high rate capability. Mater. Chem. Phys. 2014, 147, 141–146. [Google Scholar] [CrossRef]
  21. Fan, Z.; Yan, J.; Wei, T.; Zhi, L.; Ning, G.; Li, T.; Wei, F. Asymmetric Supercapacitors Based on Graphene/MnO2 and Activated Carbon Nanofiber Electrodes with High Power and Energy Density. Adv. Funct. Mater. 2011, 21, 2366–2375. [Google Scholar] [CrossRef]
  22. Liu, Y.; Shi, K.; Zhitomirsky, I. Asymmetric supercapacitor, based on composite MnO2-graphene and N-doped activated carbon coated carbon nanotube electrodes. Electrochim. Acta 2017, 233, 142–150. [Google Scholar] [CrossRef]
  23. Le, Q.J.; Huang, M.; Wang, T.; Liu, X.Y.; Sun, L.; Guo, X.L.; Jiang, D.B.; Wang, J.; Dong, F.; Zhang, Y.X. Biotemplate derived three dimensional nitrogen doped graphene@MnO2 as bifunctional material for supercapacitor and oxygen reduction reaction catalyst. J. Colloid Interface Sci. 2019, 544, 155–163. [Google Scholar] [CrossRef] [PubMed]
  24. Jin, Y.; Chen, H.; Chen, M.; Liu, N.; Li, Q. Graphene-patched CNT/MnO2 nanocomposite papers for the electrode of high-performance flexible asymmetric supercapacitors. ACS Appl. Mater. Interfaces 2013, 5, 3408–3416. [Google Scholar] [CrossRef]
  25. Zhang, X.; Yu, P.; Zhang, H.; Zhang, D.; Sun, X.; Ma, Y. Rapid hydrothermal synthesis of hierarchical nanostructures assembled from ultrathin birnessite-type MnO2 nanosheets for supercapacitor applications. Electrochim. Acta 2013, 89, 523–529. [Google Scholar] [CrossRef]
  26. Shinde, P.A.; Lokhande, V.C.; Ji, T.; Lokhande, C.D. Facile synthesis of hierarchical mesoporous weirds-like morphological MnO2 thin films on carbon cloth for high performance supercapacitor application. J. Colloid Interface Sci. 2017, 498, 202–209. [Google Scholar] [CrossRef]
  27. Liu, J.; Zhang, L.; Wu, H.B.; Lin, J.; Shen, Z.; Lou, X.W. High-performance flexible asymmetric supercapacitors based on a new graphene foam/carbon nanotube hybrid film. Energy Environ. Sci. 2014, 7, 3709–3719. [Google Scholar] [CrossRef]
  28. Nardecchia, S.; Carriazo, D.; Ferrer, M.L.; Gutierrez, M.C.; del Monte, F. Three dimensional macroporous architectures and aerogels built of carbon nanotubes and/or graphene: Synthesis and applications. Chem. Soc. Rev. 2013, 42, 794–830. [Google Scholar] [CrossRef]
  29. Liu, F.; Wang, Z.; Zhang, H.; Jin, L.; Chu, X.; Gu, B.; Huang, H.; Yang, W. Nitrogen, oxygen and sulfur co-doped hierarchical porous carbons toward high-performance supercapacitors by direct pyrolysis of kraft lignin. Carbon 2019, 149, 105–116. [Google Scholar] [CrossRef]
  30. Ping, Y.; Liu, Z.; Li, J.; Han, J.; Yang, Y.; Xiong, B.; Fang, P.; He, C. Boosting the performance of supercapacitors based hierarchically porous carbon from natural Juncus effuses by incorporation of MnO2. J. Alloys Compd. 2019, 805, 822–830. [Google Scholar] [CrossRef]
  31. Ramezani, M.; Fathi, M.; Mahboubi, F. Facile synthesis of ternary MnO2/graphene nanosheets/carbon nanotubes composites with high rate capability for supercapacitor applications. Electrochim. Acta 2015, 174, 345–355. [Google Scholar] [CrossRef]
  32. Li, W.; Xu, H.; Cui, M.; Zhao, J.; Liu, F.; Liu, T. Synthesis of sulfonated graphene/carbon nanotubes/manganese dioxide composite with high electrochemical properties. Ionics 2018, 25, 999–1006. [Google Scholar] [CrossRef]
  33. Guan, L.; Pan, L.; Peng, T.; Gao, C.; Zhao, W.; Yang, Z.; Hu, H.; Wu, M. Synthesis of Biomass-Derived Nitrogen-Doped Porous Carbon Nanosheests for High-Performance Supercapacitors. Acs Sustain. Chem. Eng. 2019, 7, 8405–8412. [Google Scholar] [CrossRef]
  34. Niu, L.; Shen, C.; Yan, L.; Zhang, J.; Lin, Y.; Gong, Y.; Li, C.; Sun, C.Q.; Xu, S. Waste bones derived nitrogen-doped carbon with high micropore ratio towards supercapacitor applications. J. Colloid Interface Sci. 2019, 547, 92–101. [Google Scholar] [CrossRef]
  35. Feng, H.; Hu, H.; Dong, H.; Xiao, Y.; Cai, Y.; Lei, B.; Liu, Y.; Zheng, M. Hierarchical structured carbon derived from bagasse wastes: A simple and efficient synthesis route and its improved electrochemical properties for high-performance supercapacitors. J. Power Sources 2016, 302, 164–173. [Google Scholar] [CrossRef]
  36. Kim, C.; Zhu, C.; Aoki, Y.; Habazaki, H. Heteroatom-doped porous carbon with tunable pore structure and high specific surface area for high performance supercapacitors. Electrochim. Acta 2019, 314, 173–187. [Google Scholar] [CrossRef]
  37. Liu, F.; Gao, Y.; Zhang, C.; Huang, H.; Yan, C.; Chu, X.; Xu, Z.; Wang, Z.; Zhang, H.; Xiao, X.; et al. Highly microporous carbon with nitrogen-doping derived from natural biowaste for high-performance flexible solid-state supercapacitor. J. Colloid Interface Sci. 2019, 548, 322–332. [Google Scholar] [CrossRef]
  38. Aguelmous, A.; El Fels, L.; Souabi, S.; Zamama, M.; Hafidi, M. The fate of total petroleum hydrocarbons during oily sludge composting: A critical review. Rev. Environ. Sci. Bio-Technol. 2019, 18, 473–493. [Google Scholar] [CrossRef]
  39. Li, X.; Liu, K.; Liu, Z.; Wang, Z.; Li, B.; Zhang, D. Hierarchical porous carbon from hazardous waste oily sludge for all-solid-state flexible supercapacitor. Electrochim. Acta 2017, 240, 43–52. [Google Scholar] [CrossRef]
  40. Kappenstein, C.; Pirault-Roy, L.; Guerin, M.; Wahdan, T.; Ali, A.A.; Al-Sagheer, F.A.; Zaki, M.I. Monopropellant decomposition catalysts. V. Thermal decomposition and reduction of permanganates as models for the preparation of supported MnOx catalysts. Appl. Catal. a-Gen. 2002, 234, 145–153. [Google Scholar] [CrossRef]
  41. Yan, J.; Fan, Z.; Wei, T.; Qian, W.; Zhang, M.; Wei, F. Fast and reversible surface redox reaction of graphene–MnO2 composites as supercapacitor electrodes. Carbon 2010, 48, 3825–3833. [Google Scholar] [CrossRef]
  42. Shen, H.; Kong, X.; Zhang, P.; Song, X.; Wang, H.; Zhang, Y. In-situ hydrothermal synthesis of δ-MnO2/soybean pod carbon and its high performance application on supercapacitor. J. Alloys Compd. 2021, 853, 157357. [Google Scholar] [CrossRef]
  43. Gao, Z.; Zhang, L.; Chang, J.; Wang, Z.; Wu, D.; Xu, F.; Guo, Y.; Jiang, K. Catalytic electrode-redox electrolyte supercapacitor system with enhanced capacitive performance. Chem. Eng. J. 2018, 335, 590–599. [Google Scholar] [CrossRef]
  44. Saleh, T.A.; Agarwal, S.; Gupta, V.K. Synthesis of MWCNT/MnO2 and their application for simultaneous oxidation of arsenite and sorption of arsenate. Appl. Catal. B Environ. 2011, 106, 46–53. [Google Scholar] [CrossRef]
  45. Wang, J.-W.; Chen, Y.; Chen, B.-Z. A Synthesis Method of MnO2/Activated Carbon Composite for Electrochemical Supercapacitors. J. Electrochem. Soc. 2015, 162, A1654–A1661. [Google Scholar] [CrossRef]
  46. Fischer, A.E.; Pettigrew, K.A.; Rolison, D.R.; Stroud, R.M.; Long, J.W. Incorporation of homogeneous, nanoscale MnO2 within ultraporous carbon structures via self-limiting electroless deposition: Implications for electrochemical capacitors. Nano Lett. 2007, 7, 281–286. [Google Scholar] [CrossRef]
  47. Ojha, G.P.; Pant, B.; Muthurasu, A.; Chae, S.-H.; Park, S.-J.; Kim, T.; Kim, H.-Y. Three-dimensionally assembled manganese oxide ultrathin nanowires: Prospective electrode material for asymmetric supercapacitors. Energy 2019, 188, 116066. [Google Scholar] [CrossRef]
  48. Ma, C.; Cao, E.; Li, J.; Fan, Q.; Wu, L.; Song, Y.; Shi, J. Synthesis of mesoporous ribbon-shaped graphitic carbon nanofibers with superior performance as efficient supercapacitor electrodes. Electrochim. Acta 2018, 292, 364–373. [Google Scholar] [CrossRef]
  49. Okhay, O.; Tkach, A. Graphene/Reduced Graphene Oxide-Carbon Nanotubes Composite Electrodes: From Capacitive to Battery-Type Behaviour. Nanomaterials 2021, 11, 5. [Google Scholar] [CrossRef]
  50. Gogotsi, Y.; Penner, R.M. Energy Storage in Nanomaterials—Capacitive, Pseudocapacitive, or Battery-like? ACS nano 2018, 12, 2081–2083. [Google Scholar] [CrossRef] [Green Version]
  51. Li, C.; Dong, X.; Zhang, Y.; Hu, J.; Liu, W.; Cui, X.; Hao, A. MnOx nanosheets anchored on a bio-derived porous carbon framework for high-performance asymmetric supercapacitors. Appl. Surf. Sci. 2020, 527, 146842. [Google Scholar] [CrossRef]
  52. Li, X.; Wang, Z.; Guo, L.; Han, D.; Li, B.; Gong, Z. Manganese oxide/hierarchical porous carbon nanocomposite from oily sludge for high-performance asymmetric supercapacitors. Electrochim. Acta 2018, 265, 71–77. [Google Scholar] [CrossRef]
  53. Li, M.; Yu, J.; Wang, X.; Yang, Z. 3D porous MnO2@carbon nanosheet synthesized from rambutan peel for high-performing supercapacitor electrodes materials. Appl. Surf. Sci. 2020, 530, 147230. [Google Scholar] [CrossRef]
  54. Zheng, X.; Liu, X.; Yang, X.; Wang, Y.; Liu, X.; Fu, A.; Guo, P.; Li, H. 3D interpenetrating networks of MnO2/Carbon-CNTs composites derived from ZIF-67 MOF and their application to supercapacitors. Colloids Surf. Physicochem. Eng. Aspects 2021, 623, 126686. [Google Scholar] [CrossRef]
  55. Tang, C.; Zhao, K.; Tang, Y.; Li, F.; Meng, Q. Forest-like carbon foam templated rGO/CNTs/MnO2 electrode for high-performance supercapacitor. Electrochim. Acta 2021, 375, 137960. [Google Scholar] [CrossRef]
  56. Feng, R.; Li, M.; Wang, Y.; Lin, J.; Zhu, K.; Wang, J.; Wang, C.; Chu, P.K. High-performance multi-dimensional nitrogen-doped N+MnO2@TiC/C electrodes for supercapacitors. Electrochim. Acta 2021, 370, 137716. [Google Scholar] [CrossRef]
  57. Gueon, D.; Moon, J.H. MnO2 Nanoflake-Shelled Carbon Nanotube Particles for High-Performance Supercapacitors. ACS Sustain. Chem. Eng. 2017, 5, 2445–2453. [Google Scholar] [CrossRef]
  58. Ren, C.; Yan, Y.; Sun, B.; Gu, B.; Chou, T.W. Wet-spinning assembly and in situ electrodeposition of carbon nanotube-based composite fibers for high energy density wire-shaped asymmetric supercapacitor. J. Colloid Interface Sci. 2020, 569, 298–306. [Google Scholar] [CrossRef]
  59. Sun, L.; Li, N.; Zhang, S.; Yu, X.; Liu, C.; Zhou, Y.; Han, S.; Wang, W.; Wang, Z. Nitrogen-containing porous carbon/alpha-MnO2 nanowires composite electrode towards supercapacitor applications. J. Alloys Compd. 2019, 789, 910–918. [Google Scholar] [CrossRef]
  60. Yang, M.; Kim, D.S.; Hong, S.B.; Sim, J.W.; Kim, J.; Kim, S.S.; Choi, B.G. MnO2 Nanowire/Biomass-Derived Carbon from Hemp Stem for High-Performance Supercapacitors. Langmuir 2017, 33, 5140–5147. [Google Scholar] [CrossRef]
  61. Zhao, Y.; Ran, W.; He, J.; Huang, Y.; Liu, Z.; Liu, W.; Tang, Y.; Zhang, L.; Gao, D.; Gao, F. High-performance asymmetric supercapacitors based on multilayer MnO2 /graphene oxide nanoflakes and hierarchical porous carbon with enhanced cycling stability. Small 2015, 11, 1310–1319. [Google Scholar] [CrossRef]
  62. Jiang, S.; Yang, B.; Lu, Y.; Xia, R.; Yu, T.; Gao, M. An aqueous symmetrical supercapacitor with high bulk pseudocapacitance induced by phase transformation of MnO2. J. Alloys Compd. 2021, 876, 160148. [Google Scholar] [CrossRef]
  63. Phattharasupakun, N.; Wutthiprom, J.; Chiochan, P.; Suktha, P.; Suksomboon, M.; Kalasina, S.; Sawangphruk, M. Turning conductive carbon nanospheres into nanosheets for high-performance supercapacitors of MnO2 nanorods. Chem. Commun. 2016, 52, 2585–2588. [Google Scholar] [CrossRef]
  64. Chavhan, M.P.; Sethi, S.R.; Ganguly, S. Vertically aligned MnO2 nanosheet electrode of controllable mass loading, counter to nanoparticulate carbon film electrode for use in supercapacitor. J. Energy Storage 2020, 32, 101851. [Google Scholar] [CrossRef]
Figure 1. Schematic procedure for the fabrication of MPC electrode materials.
Figure 1. Schematic procedure for the fabrication of MPC electrode materials.
Nanomaterials 11 02715 g001
Figure 2. SEM images of the MPC-A-1 (a), MPC-A-3 (b), MPC-A-5 (c), and MPC-3 (d).
Figure 2. SEM images of the MPC-A-1 (a), MPC-A-3 (b), MPC-A-5 (c), and MPC-3 (d).
Nanomaterials 11 02715 g002
Figure 3. TEM images of the MPC-A-3 composite material (ac) with corresponding elemental mapping images of C (d), Mn (e), and O (f).
Figure 3. TEM images of the MPC-A-3 composite material (ac) with corresponding elemental mapping images of C (d), Mn (e), and O (f).
Nanomaterials 11 02715 g003
Figure 4. XRD patterns (a), TGA analyses (b), and N2 adsorption–desorption isotherms (c) of the MPC materials, and the FTIR spectra (d) of HPC, HPC-A, and MPC-A-3.
Figure 4. XRD patterns (a), TGA analyses (b), and N2 adsorption–desorption isotherms (c) of the MPC materials, and the FTIR spectra (d) of HPC, HPC-A, and MPC-A-3.
Nanomaterials 11 02715 g004
Figure 5. Scheme of the possible preparation mechanism of MPC materials.
Figure 5. Scheme of the possible preparation mechanism of MPC materials.
Nanomaterials 11 02715 g005
Figure 6. CV curves of different MPC material sat a scan rate of 20 mV s−1 (a). CV curves of MPC-A-3 at different scan rates (b). GCD curves of MPC-A-3 at different charge/discharge current densities (c). Rate capability of different MPCs material (d). The charge/discharge cycling ability (e) of MPC-A-3, the inset is the GCD curves after 3000 cycles. Nyquist plots of the different MPCs material (f), the inset is the enlarged image of the high frequency zone.
Figure 6. CV curves of different MPC material sat a scan rate of 20 mV s−1 (a). CV curves of MPC-A-3 at different scan rates (b). GCD curves of MPC-A-3 at different charge/discharge current densities (c). Rate capability of different MPCs material (d). The charge/discharge cycling ability (e) of MPC-A-3, the inset is the GCD curves after 3000 cycles. Nyquist plots of the different MPCs material (f), the inset is the enlarged image of the high frequency zone.
Nanomaterials 11 02715 g006
Figure 7. (a) Current vs. scan rate of the MPC-A-3 electrode at a potential of 0.5 V. (b) CV data at a potential of 0.5 V for the calculation of k1, (c) the storage contributions from the pseudo capacitance and EDLC at a scan rate of 20 mV s −1. (d) Ratio of pseudo capacitance contributions in MPC-A-3 at different scan rates.
Figure 7. (a) Current vs. scan rate of the MPC-A-3 electrode at a potential of 0.5 V. (b) CV data at a potential of 0.5 V for the calculation of k1, (c) the storage contributions from the pseudo capacitance and EDLC at a scan rate of 20 mV s −1. (d) Ratio of pseudo capacitance contributions in MPC-A-3 at different scan rates.
Nanomaterials 11 02715 g007
Figure 8. CV curves of the ASC devices in different potential windows (a) at 20 mV s−1 and CV curves at various scan rates in the working potential window of 2.0 V (b). GCD curves at different current densities from 0.5 to 20 A g−1 (c) and Cs calculated from the GCD curves (d). Cycling durability and coulombic efficiency of ASC at a current density of 3 A g−1 (e). Ragone plot of the as-assembled ASC devices (f).
Figure 8. CV curves of the ASC devices in different potential windows (a) at 20 mV s−1 and CV curves at various scan rates in the working potential window of 2.0 V (b). GCD curves at different current densities from 0.5 to 20 A g−1 (c) and Cs calculated from the GCD curves (d). Cycling durability and coulombic efficiency of ASC at a current density of 3 A g−1 (e). Ragone plot of the as-assembled ASC devices (f).
Nanomaterials 11 02715 g008
Table 1. Comparison of the energy density with other reported ASC.
Table 1. Comparison of the energy density with other reported ASC.
Related ASC DevicesCurrent CollectorPotential Window (V)ElectrolyteSpecific Capacitance (F g−1)Energy Density
(W h kg−1)
Reference
MPC//HPCNi foam(0–2)1 M Na2SO4105.6 at 0.5 A g−158.67 at 498.8 W kg−1This work
MnO2@R//porous carbonNi foam(0–2)1 M Na2SO425.8 at 0.5 A g−19.2 at 1283.7 W kg−1[53]
MnO2/CNTs//ACNi sheet(0–1.6)2 M KOH-11.95 at 398.3 W kg−1[54]
MnO2/rGO/CNTs//ACCarbon foam(0–1.8)1 M Na2SO454.4 at 0.5 A g−141.6 at 513.7 W kg−1[55]
MnO2@TiC/C//AC-(0–1.5)1 M Na2SO476.5 at 0.5 A g−123.9 at 540 W kg−1[56]
MnO2/CNTs//ACAl foil(0–1.8)1 M Na2SO460 at 0.25 A g−127 at 225 W kg−1[57]
MnO2/CNT//PPy/CNT-(0–1.8)1 M Na2SO4-4.82 at 110 W kg−1[58]
MnO2/HNPC//HNPCNi foam(0–1.8)2 M Ca(NO3)266.9 at 1 A g−130.1 at 900 W kg−1[59]
MnO2/HC//HCTi foil(0–2)1 M Na2SO4-33.3 at 7800 W kg−1[60]
MnO2/GO//HPCNi foam(0–2)1 M Na2SO484 at 0.1 A g−146.7 at 100 W kg−1[61]
α-MnO2//α-MnO2Ni foam(0–2)1 M KOH139.9 at 0.5 A g−138.9 at 870.3 W kg−1[62]
MnO2/CN/PVDF// MnO2/CN/PVDFCarbon fiber paper(0–2)0.5 M Na2SO4-64.39 at 3870 W kg−1[63]
MnO2//CarbonNi foam(0–2)0.5 M K2SO4115 at 1 mA cm−262.3 at 5200 W kg−1[64]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, X.; Han, D.; Gong, Z.; Wang, Z. Nest-Like MnO2 Nanowire/Hierarchical Porous Carbon Composite for High-Performance Supercapacitor from Oily Sludge. Nanomaterials 2021, 11, 2715. https://doi.org/10.3390/nano11102715

AMA Style

Li X, Han D, Gong Z, Wang Z. Nest-Like MnO2 Nanowire/Hierarchical Porous Carbon Composite for High-Performance Supercapacitor from Oily Sludge. Nanomaterials. 2021; 11(10):2715. https://doi.org/10.3390/nano11102715

Chicago/Turabian Style

Li, Xiaoyu, Dong Han, Zhiqiang Gong, and Zhenbo Wang. 2021. "Nest-Like MnO2 Nanowire/Hierarchical Porous Carbon Composite for High-Performance Supercapacitor from Oily Sludge" Nanomaterials 11, no. 10: 2715. https://doi.org/10.3390/nano11102715

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop