Next Article in Journal
Bright Silicon Carbide Single-Photon Emitting Diodes at Low Temperatures: Toward Quantum Photonics Applications
Next Article in Special Issue
Control of Ni/β-Ga2O3 Vertical Schottky Diode Output Parameters at Forward Bias by Insertion of a Graphene Layer
Previous Article in Journal
Computational Study of the Curvature-Promoted Anchoring of Transition Metals for Water Splitting
Previous Article in Special Issue
Functionalized Activated Carbon Derived from Palm Kernel Shells for the Treatment of Simulated Heavy Metal-Contaminated Water
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nelumbo nucifera Seed–Derived Nitrogen-Doped Hierarchically Porous Carbons as Electrode Materials for High-Performance Supercapacitors

by
Lok Kumar Shrestha
1,*,
Rekha Goswami Shrestha
1,*,
Rashma Chaudhary
2,
Raja Ram Pradhananga
2,
Birendra Man Tamrakar
3,
Timila Shrestha
2,
Subrata Maji
1,
Ram Lal Shrestha
2,* and
Katsuhiko Ariga
1,4
1
International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), 1-1 Namiki, Tsukuba 305-0044, Ibaraki, Japan
2
Amrit Campus, Tribhuvan University, Kathmandu 44613, Nepal
3
Tri-Chandra Multiple Campus, Tribhuvan University, Kathmandu 44600, Nepal
4
Department of Advanced Materials Science, Graduate School of Frontier Sciences, The University of Tokyo, 5-1-5 Kashiwanoha, Kashiwa 277-8561, Chiba, Japan
*
Authors to whom correspondence should be addressed.
Nanomaterials 2021, 11(12), 3175; https://0-doi-org.brum.beds.ac.uk/10.3390/nano11123175
Submission received: 1 November 2021 / Revised: 19 November 2021 / Accepted: 21 November 2021 / Published: 23 November 2021
(This article belongs to the Special Issue Carbon Nanostructures as Promising Future Materials)

Abstract

:
Biomass-derived activated carbon materials with hierarchically nanoporous structures containing nitrogen functionalities show excellent electrochemical performances and are explored extensively in energy storage and conversion applications. Here, we report the electrochemical supercapacitance performances of the nitrogen-doped activated carbon materials with an ultrahigh surface area prepared by the potassium hydroxide (KOH) activation of the Nelumbo nucifera (Lotus) seed in an aqueous electrolyte solution (1 M sulfuric acid: H2SO4) in a three-electrode cell. The specific surface areas and pore volumes of Lotus-seed–derived carbon materials carbonized at a different temperatures, from 600 to 1000 °C, are found in the range of 1059.6 to 2489.6 m2 g−1 and 0.819 to 2.384 cm3 g−1, respectively. The carbons are amorphous materials with a partial graphitic structure with a maximum of 3.28 atom% nitrogen content and possess hierarchically micro- and mesoporous structures. The supercapacitor electrode prepared from the best sample showed excellent electrical double-layer capacitor performance, and the electrode achieved a high specific capacitance of ca. 379.2 F g−1 at 1 A g−1 current density. Additionally, the electrode shows a high rate performance, sustaining 65.9% capacitance retention at a high current density of 50 A g−1, followed by an extraordinary long cycle life without any capacitance loss after 10,000 subsequent charging/discharging cycles. The electrochemical results demonstrate that Nelumbo nucifera seed–derived hierarchically porous carbon with nitrogen functionality would have a significant probability as an electrical double-layer capacitor electrode material for the high-performance supercapacitor applications.

Graphical Abstract

1. Introduction

Supercapacitors, or electrical double-layer capacitors (EDLCs), store charges in the form of electrical double-layers at the electrode surface by the diffusion of electrolyte ions from the electrolyte solution, are the current state-of-the-art energy-storage systems for the storage of electrochemical energy [1,2,3,4,5,6,7,8,9,10]. Recently, supercapacitors have attracted significant attention because of their enormous high power density (>400 kW kg−1), extremely rapid charging or rapid reversible adsorption/desorption of electrolyte ions at the electrode surface, extraordinary long cycling stability without any capacitance loss (>10,000), high rate performance, low-cost and easy operation, as well as environmentally friendliness [11,12,13,14,15,16,17]. However, despite the several merits supercapacitors suffer from, the low energy density (1–20 Wh kg−1) compared to the lithium-ion batteries (>160 Wh kg−1), limiting their practical applications in devices.
The energy density is proportional to the electrode material’s specific capacitance (Cs) and the potential operating window of the cell (V). To improve the energy density, one has to either increase the Cs of the electrode material or widen the potential window. The working potential window or voltage can be achieved by replacing electrolyte solution with aqueous to non-aqueous solvents. Moreover, ionic liquids can also be used, as they can be operated over a more comprehensive voltage range than water [18,19,20,21]. However, due to the safety issue, non-aqueous solvents or ionic liquids are less explored in practical supercapacitor devices, demonstrating the need to improve the electrode material’s textural properties. Electrode materials for supercapacitors require a large specific surface area and plenty of well-defined pores for the adsorption of large numbers of electrolyte ions so that the energy storage capacity could be enhanced significantly. Therefore, nanoporous materials with a subtle balance of micro- and mesopore structure and high conductivity have become promising electrode materials in supercapacitors applications. Microporosity contributes to enhancing the electrical double-layer formation, while mesoporosity promotes the electrolyte ions’ diffusion and contributes to enhancing the rate performance. Investigations have shown that pore size distribution is a critical factor in controlling the effective surface area. The materials with hierarchically porous structures consisting of both the micro and mesopore architectures display better performance as electrode materials than the other conventional materials [22,23,24,25,26].
Among the various nanoporous materials studied, activated carbon materials due to high specific surface areas and abundant pores, high electrical conductivity and heteroatom doping display the high specific capacitance [27]. Such micro- and mesoporous carbon with nitrogen functionality can accommodate more electrolyte ions and increase the performance of the electrode materials in supercapacitor applications. Several nanoporous carbon materials such as fullerenes or fullerene crystals-derived porous carbons, carbon fibers, carbon nanotubes (CNTs), reduced graphene oxides and metal–organic framework (MOF)-derived porous carbon materials have been explored as the electrode materials of supercapacitors [28,29,30,31,32,33,34,35,36,37,38]. These carbon materials show good capacitive behavior, but the limitation in the scale-up synthesis sustainability of the energy storage system is low.
Nanoporous activated carbons prepared from biomass or biopolymers are the leading supercapacitor electrode materials due to the ultrahigh surface area, large pore volumes, good electrical conductivity, high chemical and thermal stability, simple preparation method and low cost [39]. Besides, heteroatom doping in the biomass-derived carbon further improves the performance of the energy-storage applications [40]. The major component of the biomass is the lignocellulose, which upon pyrolysis at moderate temperature (~200 to 300 °C), transforms to porous biochar. The biochar can be further activated by mixing with chemical activating agents such as KOH, zinc chloride, and phosphoric acid and carbonized at higher temperatures (up to 1000 °C) to enhance the porosity and the specific surface area. Several biomass-derived nanoporous carbons have been reported to show high surface area above 2000 m2 g−1, depending on the precursor itself, carbonization temperature, chemical activating agent and mixing ratio with the activating agent [41,42,43,44,45,46]. Due to high surface area and well-developed porosity, the biomass carbons perform excellently as the electrode materials for the electrochemical supercapacitors. For example, Choerospondias axillaris (Lapsi) seed–derived carbon material exhibited a high specific surface area of ca. 2272.3 m2 g−1 with interconnected mesoporous structure, and the electrode showed excellent supercapacitance performance giving a high specific capacitance of ca. 284 F g−1 at a current density of 1 A g−1 with a high rate performance sustaining 67.7% capacitance at 20 A g−1 and long cycle stability retaining 99% capacitance after 10,000 charging/discharging cycles [47]. Activated carbons are also equally explored in the hybrid capacitor to enhance electrochemical energy storage. For example, Minakshi and co-worker [48] reported a hybrid capacitor comprising mixed transition-metal sodium phosphate/activated carbon, which exhibited a specific discharge capacitance of 45 F g−1 over 1000 cycles. In a recent study, Zhu and co-workers [49] reported nitrogen-doped (5 at%) hierarchically porous carbon from biodecomposited products with an exceptionally high surface area of 3142 m2 g−1. Due to the improved textural properties, the material showed outstanding performance as the electrode materials for supercapacitor applications giving a high specific capacity of 209 F g−1 at 0.05 A g−1. However, a capacity loss of about 9% was observed after 10,000 cycles. Similarly, Wu and co-workers [50] successfully fabricated unique porous microrods from albizia flowers, and the derived material possessed hierarchically porous architectures. The materials showed a high surface area of 2757.6 m2 g−1, and self-nitrogen doping of 1.34 wt% was found. Furthermore, the albizia flowers-derived porous microrods showed a high specific capacitance of 406 F g−1 at 0.5 A g−1. In a separate work, Chen and co-workers [51] also fabricated the self-nitrogen-doped three-dimensional (3D) nanoporous carbon materials from waste cottonseed husk with honeycomb-like interconnected hierarchical porous structures. Due to a suitable surface area of 1694.1 m2 g−1, the cottonseed husk carbon material performs excellently as the electrode material for supercapacitors. The electrode showed a high specific capacitance of 238 F g−1 at 0.5 A g−1 with a relatively good cycle stability of 91% after 5000 charging/discharging cycles. Wickramaarachchi and co-workers [52] have recently reported KOH-activated carbon material from a bio-waste, Mango seed husk. The material obtained by the carbonization at 1100 °C showed the best textural properties with a high specific surface area of 1943 m2 g−1 and an average pore volume of 0.397 cm3 g−1. As a result, the carbon exhibited a maximum capacitance of 135 F g−1 at 5 mA cm−2 with the energy density of 19 Wh kg−1 at the power density of 1077 W kg−1. We recently found that Lotus seed yields nanoporous carbon material upon activation with zinc chloride with moderate surface areas and pore volumes (1103–1316 m2 g−1 and 0.741–0.887 cm3 g−1) [53]. Furthermore, the electrode prepared from the optimal sample performs reasonably well as the electrical double-layer capacitor achieving the specific capacitance of 272.9 F g−1 at 1 A g−1 indicating the possibility of further enhancing the energy storage capacity by optimizing the surface textural properties of Lotus-seed–derived carbon materials. These examples demonstrate the importance of biomass for producing electrode materials with good textural properties and self-nitrogen-doping for enhancing the overall properties of the supercapacitors.
In this work, we synthesized self-nitrogen-doped ultrahigh surface area nanoporous carbon material with hierarchically micro- and mesoporous structures from Nelumbo nucifera–seed powder and studied its electrochemical energy storage performance, the electrode material for the supercapacitor in an aqueous electrolyte solution (1 M H2SO4) in a three-electrode cell. The fabrication method includes mixing biochar of Nelumbo nucifera seed in potassium hydroxide (KOH) and carbonization at higher temperatures from 600 to 1000 °C in an inert atmosphere of nitrogen. Due to the high surface area, well-developed porosity, nitrogen doping and partially developed graphitic carbon structure, the electrode prepared from the optimal sample showed excellent electrical double-layer capacitor performance achieving a high specific capacitance of ca. 379.2 F g−1 at 1 A g−1 current density followed by a high rate performance of 65.9% at a high current density of 50 A g−1 and extraordinary long cycle life without any significant capacitance loss after 10,000 charging/discharging cycles. These results indicate the potential of Nelumbo nucifera seed as the natural precursors for the large-scale and cost-effective production of the hierarchically porous nitrogen-doped activated carbon materials essential for the development of sustainable electrode material for high-performance supercapacitors.

2. Materials and Methods

2.1. KOH Activation of Nelumbo Nucifera Precursor

After adequately washing with Milli-Q filtered water Nelumbo nucifera (Lotus) seed was dried at 100 °C for 24 h and crushed to powder form, using a mechanical crusher. The Lotus-seed powder (100 g) was heated at 300 °C for 6 h under the air atmosphere to obtain biochar. The yield of the biochar was ca. 44%. The biochar (1 g) was then mixed with KOH pellet at a 1:1 weight ratio and ground and stored at 25 °C for 24 h. The mixture was carbonized at high temperatures, from 600 to 1000 °C, under a nitrogen gas atmosphere. The temperature ramp, hold time and nitrogen gas flow were 5 °C min–1, 3 h and 120 cc min−1, respectively. The carbonization was performed in a tube furnace (KOYO, Tokyo, Japan). After the carbonization, the obtained samples were mixed with a dilute hydrochloric acid solution (0.5 M); stirred for 3 h, using the magnetic stirrer; and washed with pure water, several times, until the solution attained a neutral pH of 7. After drying in vacuum at 80 °C for 12 h, all the samples were ground into powders and designated as LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000, where the numbers indicate the carbonization temperature. After the carbonization, the mass of material was found to be 0.26 g (LTSC_K600), 0.24 mg (LTSC_K700), 0.23 g (LTSC_K800), 0.21 g (LTSC_K900) and 0.18 g (LTSC_K1000). For comparison, Lotus seed (1 g) was directly carbonized at 800 °C without KOH and designated as LTSC_800. After the direct carbonization, 0.24 g of the LTSC_800 was obtained. The KOH activation for the generation of pore structure in carbon materials is a well-known phenomenon. It includes the etching of the carbon skeleton to develop pores through gas production and washing off the potassium compounds. The KOH activation involves the following reactions (the reduction of K compounds, oxidation of C and other intermediate reactions):
6KOH + C → 2K + 2K2CO3
K2CO3 → K2O + CO
CO2 + C → 2CO
K2CO3 + 2C → 2K + 3CO
K2O + C → 2K + CO

2.2. Characterizations of Nelumbo nucifera–Derived Nanoporous Carbons

The prepared nanoporous activated carbon materials were subjected to advanced characterizations, including thermogravimetric analysis (TGA) (SII Instrument, Model Exstar 600, Tokyo, Japan), Fourier-transform infrared (FTIR) (Nicolet 4700, Thermo Electron Corporation, Waltham, MA, USA) spectroscopy, powder X-ray diffraction (XRD) (Rigaku X-ray diffractometer, RINT, Tokyo, Japan), Raman scattering (NRS-3100, JASCO, Tokyo, Japan), X-ray photoelectron spectroscopy (XPS) (Theta Probe spectrometer, Thermo Electron Co. Karlsruhe, Germany) and scanning electron microscopy (S-4800, Hitachi Co., Ltd. Tokyo, Japan). In addition, the textural properties, including specific surface area and pore-volume and pore-size distributions, were estimated by nitrogen sorption measurements (Quantachrome Autosorb-iQ2, Boynton Beach, FL, USA).

2.3. Electrochemical Studies

Using cyclic voltammetry (CV), galvanostatic charge/discharge (GCD) and electrochemical impedance spectroscopy (EIS), we studied the electrochemical performances of the Lotus-seed–derived nanoporous carbon materials as the electrode materials for supercapacitor applications. Working electrodes were prepared on the glassy carbon electrode (GCE: outer and inner diameters of 10 and 5 mm, respectively). First, carbon material was dispersed in a mixed solvent (water:ethanol = 4:1, 2 mg mL−1) and sonicated for 60 min. Next, the suspension (3 µL) aliquot was drop cast at the center of the GCE (5 mm diameter) and dried at 60 °C for 2 h for the evaporation of solvents, which gave the mass of the active material on the electrode is 6 × 10−3 mg. Finally, a Nafion solution (5% in ethanol: 5 µL) was added on top of the carbon sample on GCE as a binder and further dried under reduced pressure at 80 °C for 12 h. In addition, a platinum wire and Ag/AgCl were used as counter and reference electrodes, respectively. The EIS measurements were performed at an amplitude of 5 mV, in the frequency range of 0.01 Hz to 100 kHz. All the electrochemical measurements were carried out in a three-electrode system in an aqueous electrolyte of 1 M H2SO4 at 25 °C. CHI 660E workstation (CH Instruments, Inc. Austin, TX, USA) was used. The specific capacitance (Cs) of the electrode material was calculated from both CV (Equation (6)) and GCD curves (Equation (7)), using the following equations.
C s = 1 m × v × Δ V V 1 V 2 I ( V ) d V
where m, v, V and I represent mass of the electrode material (g), scan rate (mV s−1), operating potential window (V) and current (A), respectively.
C s = I × t d m × Δ V
where I, td, m and ΔV respectively represent the discharge current (A), the discharge time (s), the mass of the electrode material (g) and the potential window (V), respectively.

3. Results

The pyrolysis properties of the carbon source, Lotus-seed powder (Figure 1a: TGA curve), reveal the carbonization process in different stages. The first stage involves the evaporation of moisture or crystallized water in the precursor below 200 °C. The polymerization of the main components of biomass, cellulose and hemicellulose takes place in the second stage, which also involves the degradation of carbohydrates and lipids in the range of 200–500 °C, releasing the volatile gases and thus causing a significant weight loss. As a result, about 70% of the mass is lost (Figure 1a). Finally, carbon formation occurs above 500 °C, with no noticeable weight loss in the TGA curve, demonstrating that the carbonization can be carried out above this temperature. Based on the pyrolysis characteristics, we conducted KOH activation of Lotus-seed biochar by carbonizing at different temperatures, namely 600, 700, 800, 900 and 1000 °C, where the carbon yields are estimated at 26.6, 24.5, 22.9, 21.2 and 18.8%, respectively.
The FTIR spectrum shows the presence of heteroatoms (oxygen and nitrogen) functionalities in the precursor (Figure 1b). An intense FTIR peak at 3285 cm−1 corresponds to the N–H (str.), bands at 2922 and 2851 cm−1 corresponding to cellulose’s aliphatic C–H (str.). The band at 1636 cm−1 relates the O–H (def.) of adsorbed moisture water in the precursor. The FITR bands in the range of 1600–100 cm−1, commonly observed in the biomass due to cellulose and lignin, correspond to C–H (def.) and C–O (str.). After carbonization, the intensity of the FTIR bands corresponding to oxygen and nitrogen functionalities decreases significantly due to high-temperature carbonization (Supplementary Materials Figure S1). The broadband at 3446 cm−1 corresponds to adsorbed water, while a weak peak at about 1630 cm−1 can be attributed to the C=C (str.) common in the biomass-derived activated carbons [54]. The surface composition of the directly carbonized sample and the KOH-activated carbon materials was further investigated by the XPS (Figure 2). The XPS survey spectra display peaks at 284, 400 and 532 eV, which confirm carbon, oxygen and nitrogen as the main components of the prepared carbon materials (Figure 2a). Note that an increase in carbonization temperature alters the surface composition changes, and less oxygen and nitrogen functionalities are present in the samples carbonized at higher temperatures. Maximum nitrogen content of 3.4 at% was observed in LTSC_K600. The C 1s spectra of the samples could be deconvoluted into four peaks, with a peak centered at 284.4, 285.1, 286.1 and 289.4 eV, which correspond to C=C, C–N, C–C and C=O bonding states of the carbon material, respectively. The deconvoluted N 1s peaks centered at 397.9 and 400.5 eV, respectively, correspond to pyridinic-N and graphitic-N [55]. The O 1s XPS spectra with the deconvoluted peaks are shown in Supplementary Materials Figure S1b the two primary bonding states of oxygen.
Using pXRD and Raman scattering spectroscopy, we studied the structure of the prepared carbon materials. All the samples show typical XRD patterns commonly observed in amorphous carbon materials (Figure 3a). Two broad diffraction peaks observed at diffraction angles of 24 and 43° can be attributed to the (002) and (100) plans of disordered graphite-like structures of amorphous carbon often realized in biomass-derived carbon materials [54]. Raman scattering spectra support the structural characteristics of these samples. Raman spectra contain two pronounced peaks located at ~1350 and 1595 cm−1, corresponding to the D (disorder or imperfect structure) and G (graphitic structure) bands of amorphous carbons (Figure 3b) [56]. The intensity ratios of G and D bands (IG/ID) are found in the range of 0.98 to 1.03, typically observed in amorphous carbon with partial graphitic structures. The presence of defects (in the form of micropore structures) in the carbon matrix is advantageous to increase the specific surface area and hence the capacitance performance. On the other hand, graphitic carbon structure contributes to enhancing the conductivity [57].
The surface morphology of the carbon materials was investigated by SEM observations. Figure 4 shows typical SEM images of the directly carbonized (LTS_800: Figure 4a,b) and KOH-activated samples (LTSC_K600, Figure 4c,d; LTSC_K700, Figure 4e,f; LTSC_K800, Figure 4g,h; LTSC_K900, Figure 4i,j; and LTSC_K1000, Figure 4k,l). Additional SEM images are supplied in Supplementary Materials Figures S2–S7. Micron-sized irregular-shaped particles (granules) are common in all the samples. Pores are not noticeable in the low- and high-resolution SEM images of the LTS_800 sample, demonstrating the lack of surface porosity. However, abundant large-size macropores with channel-like structures are observed in the KOH-activated samples (Figure 4c,e,g,i,k and Supplementary Materials Figures S2–S7). The frameworks of these macroporous channels comprised well-developed micro- and mesopore structures due to the KOH activation, suggesting the hierarchical pore structures. Therefore, the KOH-activated samples are expected to display a high specific surface area and well-defined pore size distribution with a large pore volume advantageous in high-energy-storage supercapacitors.
Nitrogen sorption measurements were performed to study the textural properties (porosity) of the prepared carbon materials. The nitrogen uptake in the LTS_800 sample is low, and the isotherm shows Type-III sorption behavior corresponding to the nonporous materials, which supports the SEM observation that the directly carbonized sample lacks well-developed porosity. While all the KOH-activated samples show mixed Type-I/Type-IV adsorption isotherms (Figure 5a). Significant nitrogen adsorption at a low relative pressure (P/P0 < 0.1), followed by a gradual nitrogen uptake at high relative pressure with a clear hysteresis loop, demonstrates the presence of hierarchically bimodal micro- and mesopore architectures. Substantial nitrogen uptake in the low relative pressure region corresponds to micropore filling, while the hysteresis loop at high relative pressure is due to capillary condensation occurring in the mesopores [24,58,59]. Careful observation of the sorption isotherms reveals that the nitrogen uptake in the low relative pressure increases with the carbonization temperature up to 900 °C and then decreases. However, the size of the hysteresis loop monotonously increases with the temperature, suggesting that the micropore coalescence takes place at 1000 °C, leading to more mesoporous structure formation in the LTSC_K1000 sample. The pore-size-distribution profiles determined from the density functional theory (DFT) method (Figure 5b) and Barrett–Joyner–Halenda (BJH) model (Figure 5c) show prominent peaks in the micro- and mesopore region, thus further confirming the presence of hierarchically pore architectures in the KOH-activated samples. The textural properties summarized in Table 1 clearly show the crucial role of the carbonization temperature in porosity development. An increase in temperature monotonously increases the total specific surface area. However, microporosity reaches a maximum at 900 °C and then decreases, caused by the micropore coalescence, leading to the formation of mesopores, demonstrating that the LTSC_K900 sample has a much larger electrochemically accessible surface area and appropriate pore size distribution (majority micropores), enhancing the electrolyte ion adsorption and thus improving the energy storage capacity.
Encouraged by the hierarchically porous architectures, ultrahigh surface area, well-developed porosity, large pore volume and self-nitrogen doping, we explored the Lotus-seed–derived activated carbons as the electrode materials for the electrical double-layer supercapacitor applications. The CV curves recorded at a fixed scan rate of 50 mV s−1 in a three-electrode system in aqueous 1M H2SO4 (Figure 6a) agree with the results obtained from the nitrogen sorption isotherms and are well correlated to the textural properties that increase microporosity increases the total integral current. The LTSC_K900 samples show the highest current output, demonstrating the highest energy-storage capacity among the studied samples. The rectangular shapes of the CV curves with weak humps in the range of 0.2–0.45 V, followed by a quick response to the current on the reversal of the potential sweep, suggest the dominance of the electrical double-layer capacitor charge storage mechanism characteristics of the carbon materials with some contribution of pseudocapacitance, due to the presence of the nitrogen and oxygen functionalities [6,23,49,60]. The CV profiles of the electrodes prepared from all the samples reveal that the integral current increases with the scan rate, while sustaining the semi-rectangular shape of the curves (Figure 6b–d and Supplementary Materials Figure S8). Note that the weak peaks have come from the pseudocapacitive behavior of the nitrogen and oxygen functionality decrease in the carbon samples prepared at higher temperatures due to loss of nitrogen- and oxygen content (Figure 2a and Figure 4b–d and Supplementary Materials Figure S8b–d). The Cs of the electrode materials calculated by using Equation (6) shows that the directly carbonized sample, due to the lack of electrochemically accessible micropores, possesses very low specific capacitance as compared to the KOH-activated samples. The optimal sample, LTSC_K900, achieved a high Cs of ca. 434.5 F g−1 at 5 mV s−1, which can be attributed to the outstanding surface area caused due to the presence of hierarchically micro- and mesopore architecture and self-nitrogen doping. Among the KOH-activated samples, the Cs follow the order LTSC_K900 > LTSC_K800 > LTSC_K1000 > LTSC_K700 > LTSC_K600, which is in good agreement with the microporosity of the materials (Table 1). Furthermore, the electrodes prepared from the KOH-activated samples show outstanding capacitance retentions at a high scan rate, 500 mV s−1 (Figure 6f). Interestingly, the Cs retention of LTSC_K1000 (83.4%) is better than it is for LTSC_K900 (81.2%), suggesting the fast electrolyte ion diffusion through the mesopore channels, even at the high potential sweep.
Figure 7 describes the results obtained from the GCD measurements between 0 and 0.8 V, at different current densities, from 1 to 50 A g−1. Figure 7a shows the GCD curves measured at a constant current density of 1 A g−1 for LTS_800 and the KOH-activated samples. The GCD curves recorded at a high current density of 50 A g−1 are shown in Supplementary Materials Figure S9a. The symmetrical quasi-triangular shaped GCD curves with the linear decay during discharging indicate the ideal capacitive behavior of the electrode materials with well-balanced charge storage [60,61]. The discharge time follows the order of LTSC_K900 > LTSC_K800 > LTSC_K1000 > LTSC_K700 > LTSC_K600 > LTS_800, which is correlated to the micropore surface area of the materials (Table 1). A small deviation from the triangular shape with a small voltage drop is caused due to nitrogen and oxygen functional groups. Note that the discharge time of the directly carbonized sample is far less than the discharge times of the KOH-activated samples, suggesting better energy-storage capacity of the activated samples. As expected from the porosity properties, the LTSC_K900 sample has an extended discharge time indicating the highest energy storage capacity among the other activated samples studied. The GCD curves recorded at different current densities (1 to 50 A g−1) show that the quasi-triangular shape of the curve is sustained even at a high current density of 50 A g−1 (Figure 7b–d and Supplementary Materials Figure S9b–d), indicating the fast ion transfer to the electrode surface for the formation of an electrical double layer with a good rate performance and well-balanced storage of the charges. Figure 8e shows the Cs calculated by using Equation (7). Due to the electrochemically accessible high microporous surface area, the LTSC_K900 sample shows the highest specific capacitance of ca. 379.2 F g−1, while the nonporous sample obtained by direct carbonization show only 7.1 F g−1 at 1 A g−1. The rate capability of the electrodes prepared from the carbon materials obtained at higher carbonization temperatures (700–1000 °C) shows better performance than the directly carbonized sample. The capacitance retentions were ca. 63.5% (LTSC_K700), 63.3% (LTSC_K800), 65.9% (LTSC_K900) and 72.4% (LTSC_K1000) at a high current density of 50 A g−1, thus suggesting the excellent rate performance essentially required for high-performance supercapacitors. Our materials’ overall electrochemical supercapacitance performance is better than the commercial activated carbons, for which Cs is reported in the range of ~100 F g−1 [52,62], and comparable to or better than the performance of nanoporous activated carbon materials derived from other biomass precursors, such as Washnut, Lapsi and Jackfruit seed; corncob; bamboo; and others (Supplementary Materials Table S1).
Cycling stability is a crucial parameter to evaluate the performance of supercapacitors. Here we have studied the cycle performance of the selected electrodes prepared by using LTSC_K800, LTSC_K900 and LTSC_K1000 samples at a fixed current density of 50 A g−1. As commonly observed in the electrical double-layer capacitive materials, all the electrodes display outstanding long-cycle performance sustaining more than 99% capacitance after the successive 10,000 charging/discharging cycles (Figure 8a). The outstanding cycle performance can be attributed to the hierarchically porous architecture with micro- and mesopore structures that contributes to the fast electrolyte ion diffusion to the electrode surface [63,64].
EIS measurements reveal the charge storage mechanism and electrolyte ions’ diffusion kinetics at the electrode surface. The straight vertical lines at the low-frequency region in the Nyquist plots (negative of imaginary part versus real part of the complex impedance) are the characteristic of typical carbon materials with double-layer capacitive behavior (Figure 8b) [65]. Depending on the sample, the electrode has a different ion diffusion resistance; for example, the LTSC_K600 sample shows non-vertical lines in the low frequency, due to the relatively high ion diffusion resistance. Due to the nitrogen and oxygen functionalities, the electrode prepared by using carbon samples carbonized at the lower temperature display weak semicircular response, suggesting a minimal charge-transfer resistance and hence a high efficiency of the ion diffusion. The values of the equivalent series’ resistance (ESR: often interpreted as the sum of the bulk electrolyte resistance, the electrode resistance and contact resistance of electrode and current collector) estimated from the intersection point of the imaginary and real part at high frequency are low, indicating good conductivity of the prepared carbon materials. They are ca. 5.12, 5.10, 4.85, 4.62 and 4.61 Ω for LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000, respectively. The EIS results demonstrate that the differences in the energy-storage capacity of the KOH-activated samples are caused mainly due to the different porosity properties (surface area and pore volumes).

4. Conclusions

In conclusion, we studied the performance of the self-nitrogen-doped nanoporous carbon materials prepared by the KOH activation of the biochar of the Lotus-seed powder by measuring the electrochemical supercapacitance in an aqueous electrolyte (1 M H2SO4) in a three-electrode system. Activated carbon materials with nanoporous bimodal pore structures comprising micro- and mesopores and self-nitrogen doping were prepared by the KOH activation of the biochar obtained by the pyrolysis of Lotus-seed powder at 300 °C. The Lotus-seed biochar was mixed with KOH and carbonized at different temperatures (600 to 1000 °C). Surface textural properties show the crucial role of the carbonization temperature on the porosity. An increase in temperature increases the total specific surface area. However, the microporosity reaches a maximum at 900 °C and declines due to the micropore coalescence at high temperatures. The surface area and pore volume were calculated in the range of 1059.6 to 2489.6 m2 g−1 and 0.819 to 2.384 cm3 g−1, respectively. The amorphous carbons materials with partial graphitic structure contained a maximum of 3.28 atom% nitrogen content and possessed hierarchically micro- and mesoporous structures. The electrode prepared from the optimal sample performed excellently as the electrical double-layer capacitive material. The electrode achieved a high specific capacitance of ca. 379.2 F g−1 at 1 A g−1 current density, excellent rate performance sustaining 65.9% capacitance retention at 50 A g−1 and extraordinary long cycle life with only 0.3% capacity loss after 10,000 subsequent charging/discharging cycles. Thus, the electrochemical results demonstrate that Nelumbo nucifera seed, an agro-waste, exemplifies a novel yet low-cost precursor for the large-scale production of hierarchically porous carbon materials with self-nitrogen doping, having significant potential as the electrical double-layer capacitor electrode material in the applications of high-performance energy-storage supercapacitors.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/nano11123175/s1. Figure S1: (a) FTIR spectra, and (b) XPS O 1s spectra with the deconvoluted peaks of the directly carbonized sample, LTS_800, and KOH activated samples; LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900, and LTSC_K1000. Figure S2: Additional SEM images of LTS_800. Figure S3: Additional SEM images of LTSC_K600. Figure S4: Additional SEM images of LTSC_K700. Figure S5: Additional SEM images of LTSC_K800. Figure S6: Additional SEM images of LTSC_K900. Figure S7: Additional SEM images of LTSC_K1000. Figure S8: (a) The CV curves of all the samples at a fixed scan rate of 5 mV s−1 recorded at 25 °C, and the CV curves vs. scan rates for (b) LTSC_K600, (c) LTSC_K700, and (d) LTSC_K1000 systems. Figure S9: (a) The GCD curves at a constant current density of 50 A g−1, and the GCD curves vs. current density for (b) LTSC_K600, (c) LTSC_K700, and (d) LTSC_K1000. Table S1: Comparison of the electrochemical supercapacitance of the KOH activated Nelumbo nucifera (Lotus) seed carbon materials with activated carbon materials derived from other biomass.

Author Contributions

Conceptualization, L.K.S., R.G.S., R.R.P. and R.L.S.; methodology, L.K.S., R.G.S., R.C., B.M.T., T.S. and R.L.S.; formal analysis, L.K.S., R.R.P. and R.G.S.; investigation, L.K.S., R.G.S., R.C., B.M.T., T.S., R.L.S. and S.M.; data curation, R.C., T.S., B.M.T. and S.M.; writing—original draft preparation, L.K.S. and R.G.S.; writing—review and editing, L.K.S. and K.A.; supervision, L.K.S.; project administration, L.K.S.; funding acquisition, L.K.S. and K.A. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partially supported by JSPS KAKENHI grant numbers JP20H00392, JP20H00316, JP20K05590 and JP21H04685.

Data Availability Statement

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

Acknowledgments

Rashma Chaudhary and Timila Shrestha are thankful to the International Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS) for the internship award, from 17 February to 12 July 2020 (R.C.) and from 18 August to 14 September 2019 and 6 January to 28 January 2020 (T.S.).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Simon, P.; Gogotsi, Y. Perspectives for Electrochemical Capacitors and Related Devices. Nat. Mater. 2020, 19, 1151–1163. [Google Scholar] [CrossRef]
  2. Jakubec, P.; Bartusek, S.; Dvořáček, J.J.; Šedajová, V.; Kupak, V.; Otyepka, M. Flax-Derived Carbon: A Highly Durable Electrode Material for Electrochemical Double-Layer Supercapacitors. Nanomaterials 2021, 11, 2229. [Google Scholar] [CrossRef]
  3. Zhou, Y.; Qi, H.; Yang, J.; Bo, Z.; Huang, F.; Islam, M.S.; Lu, X.; Dai, L.; Amal, R.; Wang, C.H.; et al. Two-Birds-One-Stone: Multifunctional Supercapacitors Beyond Traditional Energy Storage. Energy Environ. Sci. 2021, 14, 1854–1896. [Google Scholar] [CrossRef]
  4. Arena, A.; Branca, C.; Ciofi, C.; D’Angelo, G.; Romano, V.; Scandurra, G. Polypyrrole and Graphene Nanoplatelets Inks as Electrodes for Flexible Solid-State Supercapacitor. Nanomaterials 2021, 11, 2589. [Google Scholar] [CrossRef] [PubMed]
  5. Mirabedini, A.; Lu, Z.; Mostafavian, S.; Foroughi, J. Triaxial Carbon Nanotube/Conducting Polymer Wet-Spun Fibers Supercapacitors for Wearable Electronics. Nanomaterials 2021, 11, 3. [Google Scholar] [CrossRef]
  6. Wang, T.; Hu, S.; Wu, D.; Zhao, W.; Yu, W.; Wang, M.; Xu, J.; Zhang, J. Boosting the Capacity of Biomass-Based Supercapacitors using Carbon Materials of Wood Derivatives and Redox Molecules from Plant. J. Mater. Chem. A 2021, 9, 11839–11852. [Google Scholar] [CrossRef]
  7. Huang, T.-F.; Hsieh, T.-H.; Tseng, F.-S.; Wang, L.-Y.; Yang, C.-C.; Yang, C.-C. High-Mass Loading Hierarchically Porous Activated Carbon Electrode for Pouch-Type Supercapacitors with Propylene Carbonate-Based Electrolyte. Nanomaterials 2021, 11, 785. [Google Scholar] [CrossRef]
  8. Tiwari, S.K.; Thakur, A.K.; Adhikari, A.D.; Zhu, Y.; Wang, N. Current Research of Graphene-Based Nanocomposites and Their Application for Supercapacitors. Nanomaterials 2020, 10, 2046. [Google Scholar] [CrossRef] [PubMed]
  9. Shrestha, R.G.; Maji, S.; Shrestha, L.K.; Ariga, K. Nanoarchitectonics of Nanoporous Carbon Materials in Supercapacitors Applications. Nanomaterials 2020, 10, 639. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, F.; Wu, X.; Yuan, X.; Liu, Z.; Zhang, Y.; Fu, L.; Zhu, Y.; Zhou, Q.; Wu, Y.; Huang, W. Latest Advances in Supercapacitors: From New Electrode Materials to Novel Device Designs. Chem. Soc. Rev. 2017, 46, 6816–6854. [Google Scholar] [CrossRef]
  11. Tagliaferri, S.; Nagaraju, G.; Panagiotopoulos, A.; Och, M.; Cheng, G.; Iacoviello, F.; Mattevi, C. Aqueous Inks of Pristine Graphene for 3D Printed Microsupercapacitors with High Capacitance. ACS Nano 2021, 15, 15342–15353. [Google Scholar] [CrossRef]
  12. Fite, M.C.; Rao, J.-Y.; Imae, T. Effect of External Magnetic Field on Hybrid Supercapacitors of Nitrogen-Doped Graphene with Magnetic Metal Oxides. Bull. Chem. Soc. Jpn. 2020, 93, 1139–1149. [Google Scholar] [CrossRef]
  13. Mao, W.; Yue, W.; Xu, Z.; Chang, S.; Hu, Q.; Pei, F.; Huang, Z.; Zhang, J.; Li, D.; Liu, G.; et al. Development of a Synergistic Activation Strategy for the Pilot-Scale Construction of Hierarchical Porous Graphitic Carbon for Energy Storage Applications. ACS Nano 2020, 14, 4741–4754. [Google Scholar] [CrossRef]
  14. Chen, Q.; Tan, X.; Liu, Y.; Liu, S.; Li, M.; Gu, Y.; Zhang, P.; Ye, S.; Yang, Z.; Yang, Y. Biomass-Derived Porous Graphitic Carbon Materials for Energy and Environmental Applications. J. Mater. Chem. A 2020, 8, 5773–5811. [Google Scholar] [CrossRef]
  15. Singh, G.; Bahadur, R.; Ruban, A.M.; Davidraj, J.M.; Su, D.; Vinu, A. Synthesis of Functionalized Nanoporous Biocarbons with High Surface Area for CO2 Capture and Supercapacitor Applications. Green Chem. 2021, 23, 5571–5583. [Google Scholar] [CrossRef]
  16. Joseph, S.; Saianand, G.; Benzigar, M.R.; Ramadass, K.; Sing, G.; Gopalan, A.-I.; Yang, J.H.; Mori, T.; Al-Muhtaseb, A.H.; Yi, J.; et al. Recent Advances in Functionalized Nanoporous Carbons Derived from Waste Resources and Their Applications in Energy and Environment. Adv. Sustain. Syst. 2021, 5, 2000169. [Google Scholar] [CrossRef]
  17. Qian, Q.; Lin, D.; Zhao, X.; Han, F. Vertically Oriented Grid-like Reduced Graphene Oxide for Ultrahigh Power Supercapacitor. Bull. Chem. Soc. Jpn. 2019, 48, 824–827. [Google Scholar] [CrossRef]
  18. Sahoo, S.; Krishnamoorthy, K.; Pazhamalai, P.; Mariappan, V.K.; Manoharan, S.; Kim, S.-J. High Performance Self-Charging Supercapacitors using a Porous PVDF-Ionic Liquid Electrolyte Sandwiched between Two-Dimensional Graphene Electrodes. J. Mater. Chem. A 2019, 7, 21693–21703. [Google Scholar] [CrossRef]
  19. Chatterjee, D.P.; Nandi, A.K. A Review on the Recent Advances in Hybrid Supercapacitor. J. Mater. Chem. A 2021, 9, 15880–15918. [Google Scholar] [CrossRef]
  20. Aken, K.L.V.; Deidaghi, M.; Gogotsi, Y. Formulation of Ionic-Liquid Electrolyte to Expand the Voltage Window of Supercapacitors. Angew. Chem. Int. Ed. 2015, 54, 4806–4809. [Google Scholar] [CrossRef]
  21. Pan, S.; Yao, M.; Zhang, J.; Li, B.; Xing, C.; Song, X.; Su, P.; Zhang, H. Recognition of Ionic Liquids as High-Voltage Electrolytes for Supercapacitors. Front. Chem. 2020, 8, 261. [Google Scholar] [CrossRef]
  22. Li, H.; Tao, Y.; Zheng, X.; Luo, J.; Kang, F.; Cheng, H.-M.; Yang, Q.-H. Ultra-Thick Graphene Bulk Supercapacitor Electrode for Compact Energy Storage. Energy Environ. Sci. 2016, 9, 3135–3142. [Google Scholar] [CrossRef]
  23. Dutta, S.; Bhaumik, A.; Wu, K.C.-W. Hierarchically Porous Carbon Derived from Polymers and Biomass: Effect of Interconnected Pores on Energy Applications. Energy Environ. Sci. 2014, 7, 3574–3592. [Google Scholar] [CrossRef]
  24. Benzigar, M.R.; Talapaneni, S.N.; Joseph, S.; Ramadass, K.; Singh, G.; Scaranto, J.; Ravon, U.; Al-Bahily, K.; Vinu, A. Recent Advances in Functionalized Micro and Mesoporous Carbon Materials: Synthesis and Applications. Chem. Soc. Rev. 2018, 47, 2680–2721. [Google Scholar] [CrossRef] [PubMed]
  25. Xie, L.; Su, F.; Xie, L.; Guo, X.; Wang, Z.; Kong, Q.; Sun, G.; Ahmad, A.; Li, X.; Yi, Z.; et al. Effect of Pore Structure and Doping Species on Charge Storage Mechanisms in Porous Carbon-Based Supercapacitors. Mater. Chem. Front. 2020, 4, 2610–2634. [Google Scholar] [CrossRef]
  26. Lyu, L.; Seong, K.-d.; Ko, D.; Choi, J.; Lee, C.; Hwang, T.; Cho, Y.; Jin, X.; Zhang, W.; Pang, H.; et al. Recent Development of Biomass-Derived Carbons and Composites as Electrode Materials for Supercapacitors. Mater. Chem. Front. 2019, 3, 2543–2570. [Google Scholar] [CrossRef]
  27. Jin, Y.; Tian, K.; Wei, L.; Zhang, X.; Guo, X. Hierarchical Porous Microspheres of Activated Carbon with a High Surface Area from Spores for Electrochemical Double-Layer Capacitors. J. Mater. Chem. A 2016, 4, 15968–15979. [Google Scholar] [CrossRef]
  28. Shrestha, L.K.; Shrestha, R.G.; Yamauchi, Y.; Hill, J.P.; Nishimura, T.; Miyazawa, K.; Kawai, T.; Okada, S.; Wakabayashi, K.; Ariga, K. Nanoporous Carbon Tubes from Fullerene Crystals as the π-Electron Carbon Source. Angew. Chem. Int. Ed. 2015, 54, 951–955. [Google Scholar] [CrossRef]
  29. Shrestha, L.K.; Shrestha, R.G.; Hill, J.P.; Tsuruoka, T.; Ji, Q.; Nishimura, T.; Ariga, K. Surfactant-Triggered Nanoarchitectonics of Fullerene C60 Crystals at a Liquid−Liquid Interface. Langmuir 2016, 32, 12511–12519. [Google Scholar] [CrossRef]
  30. Bairi, P.; Maji, S.; Hill, J.P.; Kim, J.H.; Ariga, K.; Shrestha, L.K. Mesoporous Carbon Cubes Derivied from Fullerene Crystals as a High Rate Performance Electrode Materials for Sueprcapacitors. J. Mater. Chem. A 2019, 7, 12654–12660. [Google Scholar] [CrossRef]
  31. Benzigar, M.R.; Joseph, S.; Ilbeygi, H.; Park, D.-H.; Sarkar, S.; Chandra, G.; Umapathy, S.; Srinivasan, S.; Talapaneni, S.; Vinu, A. Highly Crystalline Mesoporous C60 with Ordered Pores: A Class of Nanomaterials for Energy Applications. Angew. Chem. Int. Ed. 2018, 57, 569–573. [Google Scholar] [CrossRef]
  32. Benzigar, M.R.; Joseph, S.; Baskar, A.V.; Park, D.-H.; Chandra, G.; Umapathy, S.; Talapaneni, N.; Vinu, A. Ordered Mesoporous C70 with Highly Crystalline Pore Walls for Energy Applications. Adv. Funct. Mater. 2018, 28, 1803701. [Google Scholar] [CrossRef]
  33. Lu, Z.; Raad, R.; Safaei, F.; Xi, J.; Liu, Z.; Foroughi, J. Carbon Nanotube Based Fiber Supercapacitor as Wearable Energy Storage. Front. Mater. 2019, 6, 138. [Google Scholar] [CrossRef]
  34. Sengottaiyan, C.; Jayavel, R.; Shrestha, R.G.; Subramani, T.; Maji, S.; Kim, J.H.; Hill, J.P.; Ariga, K.; Shrestha, L.K. Indium Oxide/Carbon Nanotube/Reduced Graphene Oxide Ternary Nanocomposite with Enhanced Electrochemical Supercapacitance. Bull. Chem. Soc. Jpn. 2019, 92, 521–528. [Google Scholar] [CrossRef]
  35. Zhu, Y.; Murali, S.; Stoller, M.D.; Ganesh, K.J.; Cai, W.; Ferreira, P.J.; Pirkle, A.; Wallace, R.M.; Cychosz, K.A.; Thommes, M.; et al. Carbon-Based Supercapacitors Produced by Activation of Graphene. Science 2011, 332, 1537–1541. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Wang, C.; Kim, J.; Tang, J.; Na, J.; Kang, Y.-M.; Kim, M.; Lim, H.; Bando, Y.; Li, J.; Yamauchi, Y. Large-Scale Synthesis of MOF-Derived Superporous Carbon Aerogels with Extraordinary Adsorption Capacity for Organic Solvents. Angew. Chem. Int. Ed. 2020, 59, 2066–2070. [Google Scholar] [CrossRef] [PubMed]
  37. Tang, J.; Salunkhe, R.R.; Zhang, H.; Malgras, V.; Ahamad, T.; Alshehri, S.M.; Kobayashi, N.; Tominaka, S.; Ide, Y.; Kim, J.H.; et al. Bimetallic Metal-Organic Frameworks for Controlled Catalytic Graphitization of Nanoporous Carbons. Sci. Rep. 2016, 6, 30295. [Google Scholar] [CrossRef]
  38. Wang, C.; Kim, J.; Tang, J.; Kim, M.; Lim, H.; Malgras, V.; You, J.; Xu, Q.; Li, J.; Yamauchi, Y. New strategies for novel MOF-derived carbon materials based on nanoarchitectures. Chem 2020, 6, 19–40. [Google Scholar] [CrossRef]
  39. Lu, H.; Zhao, X.S. Biomass-Derived Carbon Electrode Materials for Supercapacitors. Sustain. Energy Fuels 2017, 1, 1265–1281. [Google Scholar] [CrossRef]
  40. Lin, T.; Chen, I.-W.; Liu, F.; Yang, C.; Bi, H.; Xu, F.; Huang, F. Nitrogen-Doped Mesoporous Carbon of Extraordinary Capacitance for Electrochemical Energy Storage. Science 2015, 350, 1508–1513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Lu, T.; Xu, X.; Zhang, S.; Pan, L.; Wang, Y.; Alshehri, S.M.; Ahamad, T.; Kim, M.; Na, J.; Hossain, M.d.S.A.; et al. High-Performance Capacitive Deionization by Lignocellulose-Derived Eco-Friendly Porous Carbon Materials. Bull. Chem. Soc. Jpn. 2020, 93, 1014–1019. [Google Scholar] [CrossRef]
  42. Liu, Y.; Xiao, Z.; Liu, Y.; Fan, L.-Z. Biowaste-Derived 3D Honeycomb-Like Porous Carbon with Binary-Heteroatom Doping for High-Performance Flexible Solid-State Supercapacitors. J. Mater. Chem. A 2018, 6, 160–166. [Google Scholar] [CrossRef]
  43. Yao, Y.; Feng, Q.; Huo, B.; Zhou, H.; Huang, Z.; Li, H.; Yan, Z.; Yang, X.; Kuang, Y. Facile Self-Templating Synthesis of Heteroatom-Doped 3D Porous Carbon Materials from Waste Biomass for Supercapacitors. Chem. Commun. 2020, 56, 11689–11692. [Google Scholar] [CrossRef]
  44. Pang, M.; Jiang, S.; Zhao, J.; Zhang, S.; Wang, R.; Li, N.; Liu, R.; Pan, Q.; Qu, W.; Xing, B. “Water-in-Salt” Electrolyte Enhanced High Voltage Aqueous Supercapacitor with Carbon Electrodes Derived from Biomass Waste-Ground Grain Hulls. RSC Adv. 2020, 10, 35545–35556. [Google Scholar] [CrossRef]
  45. Shrestha, R.L.; Chaudhary, R.; Shrestha, R.G.; Shrestha, T.; Maji, S.; Ariga, K.; Shrestha, L.K. Washnut Seed–Derived Ultrahigh Surface Area Nanoporous Carbons as High Rate Performance Electrode Material for Supercapacitors. Bull. Chem. Soc. Jpn. 2021, 94, 565–572. [Google Scholar] [CrossRef]
  46. Szczęśniak, B.; Phuriragpitikhon, J.; Choma, J.; Jaroniec, M. Recent Advances in the Development and Applications of Biomass-Derived Carbons with Uniform Porosity. J. Mater. Chem. A 2020, 8, 18464–18491. [Google Scholar] [CrossRef]
  47. Shrestha, L.K.; Shrestha, R.G.; Maji, S.; Pokharel, B.P.; Rajbhandari, R.; Shrestha, R.L.; Pradhananga, R.; Hill, J.P.; Ariga, K. High Surface Area Nanoporous Graphitic Carbon Materials Derived from Lapsi Seed with Enhanced Supercapacitance. Nanomaterials 2020, 10, 728. [Google Scholar] [CrossRef] [Green Version]
  48. Minakshi, M.; Meyrick, D.; Appadoo, D. Maricite (NaMn1/3Ni1/3Co1/3PO4)/Activated Carbon: Hybrid Capacitor. Energy Fuels 2013, 27, 3516–3522. [Google Scholar] [CrossRef]
  49. Zhu, Y.; Chen, M.; Zhang, Y.; Zhao, W. A Biomass-Derived Nitrogen-Doped Porous Carbon for High-Energy Supercapacitor. Carbon 2018, 140, 404–412. [Google Scholar] [CrossRef]
  50. Wu, F.; Gao, J.; Zhai, X.; Xie, M.; Sun, Y.; Kang, H.; Tian, Q.; Qiu, H. Hierarchical Porous Carbon Microrods Derived from Albizia Flower for High Performance Supercapacitors. Carbon 2019, 147, 242–251. [Google Scholar] [CrossRef]
  51. Chen, H.; Wang, G.; Chen, L.; Dai, B.; Yu, F. Three-Dimensional Honeycomb-Like Porous Carbon with Both Interconnected Hierarchical Porosity and Nitrogen Self-Doping from Cotton Seed Husk for Supercapacitor Electrode. Nanomaterials 2018, 8, 412. [Google Scholar] [CrossRef] [Green Version]
  52. Wickramaarachchi, W.A.M.K.P.; Minakshi, M.; Gao, X.; Dabare, R.; Wong, K.W. Hierarchical Porous Carbon from Mango Seed Husk for Electro-Chemical Energy Storage. Chem. Eng. J. Adv. 2021, 8, 100158. [Google Scholar] [CrossRef]
  53. Shrestha, R.L.; Chaudhary, R.; Shrestha, T.; Tamrakar, B.M.; Shrestha, R.G.; Maji, S.; Hill, J.P.; Ariga, K.; Shrestha, L.K. Nanoarchitectonics of Lotus Seed Derived Nanoporous Carbon Materials for Supercapacitor Applications. Materials 2020, 13, 5434. [Google Scholar] [CrossRef]
  54. Xu, F.; Yu, J.; Tesso, T.; Dowell, F.; Wang, D. Qualitative and Quantitative Analysis of Lignocellulosic Biomass Using Infrared Techniques: A Mini-Review. Appl. Energy 2013, 104, 801–809. [Google Scholar] [CrossRef] [Green Version]
  55. Guo, D.; Shibuya, R.; Akiba, C.; Saji, S.; Kondo, T.; Nakamura, J. Active Sites of Nitrogen-Doped Carbon Materials for Oxygen Reduction Reaction Clarified using Model Catalysts. Science 2016, 351, 361–365. [Google Scholar] [CrossRef] [PubMed]
  56. Shao, Y.; El-Kady, M.F.; Sun, J.; Li, Y.; Zhang, Q.; Zhu, M.; Wang, H.; Dunn, B.; Kaner, R.B. Design and Mechanisms of Asymmetric Supercapacitors. Chem. Rev. 2018, 118, 9233–9280. [Google Scholar] [CrossRef]
  57. Mehra, P.; Singh, C.; Cherian, I.; Giri, A.; Paul, A. Diciphering the Incredible Supercapacitor Performance Conducting Biordered Ultramicroporous Graphitic Carbon. ACS Appl. Energy Mater. 2021, 4, 4416–4427. [Google Scholar] [CrossRef]
  58. Golden, T.C.; Sircar, S. Activated Carbon Adsorbent for PSA Driers. Carbon 1990, 28, 683–690. [Google Scholar] [CrossRef]
  59. Schneider, P.; Hudec, P.; Solcova, O. Pore-Volume and Surface Area in Microporous–Mesoporous Solids. Microporous Mesoporous Mater. 2008, 115, 491–496. [Google Scholar] [CrossRef]
  60. Simon, P.; Gogotsi, Y. Materials for Electrochemical Capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [Green Version]
  61. Karnan, M.; Subramani, K.; Sudhan, N.; Ilayaraja, N.; Sathish, M. Aloe Vera Derived Activated High-Surface-Area Carbon for Flexible and High-Energy Supercapacitors. ACS Appl. Mater. Interfaces 2016, 8, 35191–35202. [Google Scholar] [CrossRef]
  62. Liu, X.; Mi, R.; Yuan, L.; Yang, F.; Fu, Z.; Wang, C.; Tang, Y. Nitrogen-Doped Multi-Scale Porous Carbo for High Voltage Aqueous Supercapacitors. Fornt. Chem. 2018, 6, 475. [Google Scholar]
  63. Jäckel, N.; Simon, P.; Gogotsi, Y.; Presser, V. Increase in Capacitance by Subnanometer Pores in Carbon. ACS Energy Lett. 2016, 1, 1262–1265. [Google Scholar] [CrossRef] [Green Version]
  64. Salanne, M.; Rotenberg, B.; Naoi, K.; Kaneko, K.; Taberna, P.L.; Grey, C.P.; Dunn, B.; Simon, P. Efficient Storage Mechanism for Building Better Supercapacitors. Nat. Energy 2016, 1, 16070. [Google Scholar] [CrossRef]
  65. Mei, B.-A.; Munteshari, O.; Lau, J.; Dunn, B.; Pilon, L. Physical Interpretations of Nyquist Plots for EDLC Electrodes and Devices. J. Phys. Chem. C 2018, 122, 194–206. [Google Scholar] [CrossRef]
Figure 1. (a) TGA curve and (b) FTIR spectrum of the carbon source, Lotus-seed powder.
Figure 1. (a) TGA curve and (b) FTIR spectrum of the carbon source, Lotus-seed powder.
Nanomaterials 11 03175 g001
Figure 2. XPS surface composition studies of the Lotus-seed–derived carbon materials. (a) XPS survey spectra, (b) XPS C 1s spectra with the deconvoluted peaks and (c) XPS N 1s spectra with deconvoluted peaks of the directly carbonized sample LTS_800 and KOH-activated LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000 samples.
Figure 2. XPS surface composition studies of the Lotus-seed–derived carbon materials. (a) XPS survey spectra, (b) XPS C 1s spectra with the deconvoluted peaks and (c) XPS N 1s spectra with deconvoluted peaks of the directly carbonized sample LTS_800 and KOH-activated LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000 samples.
Nanomaterials 11 03175 g002
Figure 3. (a) pXRD and (b) Raman scattering spectra recorded at 25 °C for Lotus-seed–derived nanoporous carbon materials; LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000.
Figure 3. (a) pXRD and (b) Raman scattering spectra recorded at 25 °C for Lotus-seed–derived nanoporous carbon materials; LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000.
Nanomaterials 11 03175 g003
Figure 4. SEM observations of the Lotus-seed–derived carbon materials: (a,b) SEM images of LTS_800, (c,d) SEM images of LTSC_K600, (e,f) SEM images of LTSC_K700, (g,h) SEM images of LTSC_K800, (i,j) SEM images of LTSC_K900 and (k,l) SEM images of LTSC_K1000.
Figure 4. SEM observations of the Lotus-seed–derived carbon materials: (a,b) SEM images of LTS_800, (c,d) SEM images of LTSC_K600, (e,f) SEM images of LTSC_K700, (g,h) SEM images of LTSC_K800, (i,j) SEM images of LTSC_K900 and (k,l) SEM images of LTSC_K1000.
Nanomaterials 11 03175 g004
Figure 5. Textural properties by nitrogen sorption measurements. (a) Nitrogen adsorption/desorption isotherms of LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000 samples; (b) pore-size-distribution profiles from the density functional theory (DFT) method; and (c) pore-size-distribution profiles obtained from the Barrett–Joyner–Halenda (BJH) model.
Figure 5. Textural properties by nitrogen sorption measurements. (a) Nitrogen adsorption/desorption isotherms of LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000 samples; (b) pore-size-distribution profiles from the density functional theory (DFT) method; and (c) pore-size-distribution profiles obtained from the Barrett–Joyner–Halenda (BJH) model.
Nanomaterials 11 03175 g005
Figure 6. Cyclic voltammetry study of the Lotus-seed–derived carbon materials. (a) The CV profiles at a fixed scan rate of 50 mV s−1, and the CV profiles at different scan rates from 5 to 500 mV s−1 for (b) LTS_800, (c) LTSC_K800 and (d) LTSC_K900 as typical examples. (e) The calculated Cs vs. scan rate. (f) The capacitance retention.
Figure 6. Cyclic voltammetry study of the Lotus-seed–derived carbon materials. (a) The CV profiles at a fixed scan rate of 50 mV s−1, and the CV profiles at different scan rates from 5 to 500 mV s−1 for (b) LTS_800, (c) LTSC_K800 and (d) LTSC_K900 as typical examples. (e) The calculated Cs vs. scan rate. (f) The capacitance retention.
Nanomaterials 11 03175 g006
Figure 7. Results obtained from the GCD measurements. (a) The GCD curves recorded at a constant current density of 1 A g−1 for LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000, and GCD curves vs. current density for (b) LTS_800, (c) LTSC_K800 and (d) LTSC_K900. (e) The Cs vs. current density. (f) The corresponding rate performance.
Figure 7. Results obtained from the GCD measurements. (a) The GCD curves recorded at a constant current density of 1 A g−1 for LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000, and GCD curves vs. current density for (b) LTS_800, (c) LTSC_K800 and (d) LTSC_K900. (e) The Cs vs. current density. (f) The corresponding rate performance.
Nanomaterials 11 03175 g007
Figure 8. (a) Cyclic stability performance of LTSC_K800, LTSC_K900 and LTSC_K1000 electrodes at 50 A g−1. (b) Nyquist plots obtained from the EIS measurements for all the samples (LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000).
Figure 8. (a) Cyclic stability performance of LTSC_K800, LTSC_K900 and LTSC_K1000 electrodes at 50 A g−1. (b) Nyquist plots obtained from the EIS measurements for all the samples (LTS_800, LTSC_K600, LTSC_K700, LTSC_K800, LTSC_K900 and LTSC_K1000).
Nanomaterials 11 03175 g008
Table 1. Textural properties of Lotus-seed–derived carbon materials carbonized at different temperatures.
Table 1. Textural properties of Lotus-seed–derived carbon materials carbonized at different temperatures.
SampleSSA (m2 g−1)Smicro (m2 g−1)Smeso (m2 g−1)Vp (cm3 g−1)Vmicro (cm3 g−1)Wp (nm)Dp (nm)
LTS_80046.118.827.30.1020.044-3.09
LTSC_K6001059.6824.5235.10.8190.4720.2853.92
LTSC_K7001878.41556.1322.31.2320.7750.2863.93
LTSC_K8002236.61891.3345.31.4991.0340.2743.91
LTSC_K9002330.11905.7424.41.7931.2060.2863.92
LTSC_K10002489.31725.6763.72.3841.4880.7053.93
SSA = total specific surface area; Smicro = micropore surface area; Vp = total pore volume; Vmicro = pore volume from micropores; Wp = average half pore width, as obtained from the DFT model; Dp = average pore diameter obtained from the BJH analysis.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shrestha, L.K.; Shrestha, R.G.; Chaudhary, R.; Pradhananga, R.R.; Tamrakar, B.M.; Shrestha, T.; Maji, S.; Shrestha, R.L.; Ariga, K. Nelumbo nucifera Seed–Derived Nitrogen-Doped Hierarchically Porous Carbons as Electrode Materials for High-Performance Supercapacitors. Nanomaterials 2021, 11, 3175. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11123175

AMA Style

Shrestha LK, Shrestha RG, Chaudhary R, Pradhananga RR, Tamrakar BM, Shrestha T, Maji S, Shrestha RL, Ariga K. Nelumbo nucifera Seed–Derived Nitrogen-Doped Hierarchically Porous Carbons as Electrode Materials for High-Performance Supercapacitors. Nanomaterials. 2021; 11(12):3175. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11123175

Chicago/Turabian Style

Shrestha, Lok Kumar, Rekha Goswami Shrestha, Rashma Chaudhary, Raja Ram Pradhananga, Birendra Man Tamrakar, Timila Shrestha, Subrata Maji, Ram Lal Shrestha, and Katsuhiko Ariga. 2021. "Nelumbo nucifera Seed–Derived Nitrogen-Doped Hierarchically Porous Carbons as Electrode Materials for High-Performance Supercapacitors" Nanomaterials 11, no. 12: 3175. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11123175

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