Next Article in Journal
Modification Strategies of g-C3N4-Based Materials for Enhanced Photoelectrocatalytic Degradation of Pollutants: A Review
Previous Article in Journal
Enhanced Photocatalytic Performances and Mechanistic Insights for Novel Ag-Bridged Dual Z-Scheme AgI/Ag3PO4/WO3 Composites
Previous Article in Special Issue
AlF3-Modified Carbon Anodes for Aluminum Electrolysis: Oxidation Resistance and Microstructural Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Physically Activated Carbons from Vitellaria paradoxa Shells for Supercapacitor Electrode Applications

by
Joshua Atta Alabi
1,2,3,4,
Neda Nazari
4,
Daniel Nframah Ampong
1,2,
Frank Ofori Agyemang
1,2,
Mark Adom-Asamoah
5,
Richard Opoku
6,
Rene Zahrhuber
7,
Christoph Unterweger
4,* and
Kwadwo Mensah-Darkwa
1,2
1
KNUST Center for Engineering Materials Research (KCEMR), Department of Materials Engineering, College of Engineering, Kwame Nkrumah University of Science and Technology, Kumasi AK-448-7139, Ghana
2
Department of Materials Engineering, College of Engineering, Kwame Nkrumah University of Science and Technology, Kumasi AK-448-7139, Ghana
3
Institute of Polymeric Materials and Testing, Johannes Kepler University, 4040 Linz, Austria
4
Wood K plus—Kompetenzzentrum Holz GmbH, 4040 Linz, Austria
5
Department of Civil Engineering, College of Engineering, Kwame Nkrumah University of Science and Technology, Kumasi AK-448-7139, Ghana
6
Department of Mechanical Engineering, College of Engineering, Kwame Nkrumah University of Science and Technology, Kumasi AK-448-7139, Ghana
7
Centre for Surface and Nanoanalytics (ZONA), Johannes Kepler University, 4040 Linz, Austria
*
Author to whom correspondence should be addressed.
Inorganics 2025, 13(7), 224; https://doi.org/10.3390/inorganics13070224
Submission received: 30 April 2025 / Revised: 13 June 2025 / Accepted: 18 June 2025 / Published: 2 July 2025

Abstract

This study investigates the processing of shea nut shells (SNSs), an abundant agricultural waste, into porous activated carbon for supercapacitor electrodes through a two-stage thermal treatment involving pyrolysis and physical activation with CO2 and steam. The aim was to develop sustainable, high-performance electrode materials while addressing waste management. Carbonization followed by activation yielded 16.5% (CO2) and 11.3% (steam) activation yields, with total yields of 4.3% and 2.9%, respectively. CO2 activation produced carbon (AC_CO2) with a specific surface area (SBET) of 1528 m2 g−1 and a total pore volume of 0.72 cm3 g−1, a graphitization degree (ID/IG = 1.0), and low charge transfer resistance (9.05 Ω), delivering a specific capacitance of 47.5 F g−1 at 0.5 A g−1, an energy density of 9.5 Wh kg−1 at 299 W kg−1, and a fast discharge time of 2.10 s, ideal for power-intensive applications. Steam activation yielded carbon (AC_H2O) with a higher specific surface area (1842 m2 g−1) and pore volume (1.57 cm3 g−1), achieving a superior specific capacitance of 102.2 F g−1 at 0.5 A g−1 and a power density of 204 W kg−1 at 9.2 Wh kg−1, suited for energy storage. AC_CO2 also exhibited exceptional cyclic stability (90% retention after 10,000 cycles). These findings demonstrate SNS-derived activated carbon as a versatile, eco-friendly material, with CO2 activation optimizing power delivery and steam activation enhancing energy capacity, offering tailored solutions for supercapacitor applications and sustainable waste utilization.

1. Introduction

Renewable energy sources, such as wind, hydroelectric power, and solar energy, have gained considerable traction as viable alternatives to fossil fuels. However, incorporating these renewable energy sources into the global energy grid presents a significant technical challenge [1]. One of the most critical issues is the unreliable nature of renewable energy production, which is also heavily affected by weather conditions [2]. Consequently, this leads to an imbalance between energy supply and demand, hence the need to look for reliable and efficient means to store excess energy. Supercapacitors (SCs) have become a promising technology in energy storage due to their unique characteristics, including rapid charge and discharge capabilities, high power density, and long cycle life. Unlike conventional batteries, SCs store energy electrostatically, enabling them to deliver rapid energy bursts and withstand numerous charge–discharge cycles with minimal degradation [3,4]. These unique properties make SCs well suited for applications that demand high power output over short durations, including electric vehicles, portable electronics, and renewable energy systems [3].
Despite their advantages, SCs are constrained by their comparatively low energy density compared to conventional batteries. This limitation restricts their application in situations that demand high energy and high-power densities. To address this challenge, extensive research has been focused on optimizing materials as electrodes in SCs [5,6], which play a vital role in defining the overall performance of these devices. Activated carbon from biomass (AC) is among the frequently utilized electrode materials because of their high specific surface area, outstanding electrical conductivity, chemical stability, and cost-effective production [7,8]. Agro-wastes such as walnut shells [9], coconut shells [10], rice husks [11], corn cobs [12], peanut shells [13], sugarcane bagasse [14], olive stones [15], banana and plantain peels [16], palm kernel shells [17], pineapple crown and peels [18], cassava peels [19], mustard seeds [20], orange peels [21], etc., have been employed in making activated carbon for electrodes in SCs.
However, several biomass sources need to be investigated when it comes to the synthesis of activated carbon. One such source is shea nut shells (SNSs), a biowaste from the shea butter production industry in Africa. With about 1,760,000 metric tons of shea nuts being processed into butter annually, around 1,000,000 metric tons of waste are produced [22]. Currently, the SNSs produced by most communities are used as waste in landfills, which serve as breeding sites for disease vectors like mosquitoes. At the same time, some industries utilize it as fuel to help sustain their production plant, contributing to carbon dioxide emissions [23]. The study of SNSs has not received much attention, hence the need to employ such agro-waste.
Several researchers have worked with SNSs as a biowaste for synthesizing activated carbon using a chemical activating method. Abdullahi et al. [24] synthesized activated carbon from SNSs and desert dates using 1.0 M of H2SO4 and investigated how contact time, pH levels, and temperature influence the efficiency of removing lead and cadmium from water. The researchers concluded that the synthesized activated carbon effectively removed lead and cadmium from contaminated water. Musah et al. [25] synthesized activated carbon with good pores of different dimensions from SNSs using H3PO4 and KOH as activating agents. They reported that synthesized carbons are effective for treating wastewater. Ampong et al. [8] also synthesized oxygen-rich interconnected hierarchical activated carbon using SNSs via a two-stage activation process using KOH, which exhibited impressive electrochemical performance, achieving a high specific capacitance of 286.6 F g−1 and enhanced rate capability. These remarkable properties stemmed from the material’s high specific surface area and O-rich content. The research works reported so far employed chemical activation when utilizing SNSs for synthesizing activated carbons, with most of the application focused on water treatment and very little attention on supercapacitor electrodes for energy storage.
Therefore, based on the above survey and using an experimental setup established in our previous work as described by Breitenbach et al. [26], activated carbons with high specific surface area are synthesized via a two-stage synthesis process by pyrolysis in a nitrogen atmosphere followed by physical activation using steam or CO2 as activating agents. The prepared carbons are comprehensively characterized and utilized as SC electrodes. With the collected data, relationships between processing, properties, and electrochemical performance are evaluated and compared to existing data from the literature.

2. Results and Discussion

2.1. Material Characterization

Physicochemical characterizations were performed on the precursor sample, as well as moisture content and thermogravimetric analysis (TGA). In Figure 1a, the TGA and DTA results of the untreated precursor sample from room temperature up to 900 °C are presented. The TGA curve shows three distinct zones: A—drying stage, B—pyrolytic stage, and C—carbonization stage. TGA of the raw SNS sample resulted in a one-step thermogram, suggesting a single dominant decomposition process. In contrast, differential thermal analysis (DTA) revealed a broad peak accompanied by shoulder peaks, indicating the presence of multiple components within the raw SNS sample [23]. The thermal analysis began with a loss of moisture at the drying stage, about 6% of moisture of the total mass of the sample at 102 °C. This agrees with the initial moisture analysis conducted on the sample using the moisture analyzer. The degradation temperature for the pyrolytic stage began around 261 °C. This temperature marks the onset of the breakdown of hemicellulose, cellulose, and other volatile compounds in the material. Maximum degradation in the pyrolytic stage was recorded at 320 °C for the onset of degradation for lignin. The carbonization in zone C marks char production with about 26% carbon yield.
Figure 1b presents the overall carbon yield produced from the two modes of physical activation. An initial carbon yield of about 26% was produced after carbonization, which agreed with the thermal analysis results. Physical activation with steam produced an activation yield of 11.3% and a total yield of 2.9%, as compared to 16.5% and 4.3% for CO2 activation. This result shows that the latter produced more carbon yield than the former, with CO2 activation producing a double carbon yield as water activation. This phenomenon is attributed to the erratic nature of the steam activation mechanism, which attacks the active sites of the carbon structure to form pores while simultaneously attacking the walls of the pores [27]. CO2 activation occurs more slowly than steam activation due to reduced diffusion rates at the active sites within the center of the pores [28]. Hence, CO2 produced a higher carbon yield than steam activation. Detailed reaction mechanisms can be found in references [27,29].
The surface functional groups in the synthesized activated carbon samples were characterized using FTIR spectroscopy, as illustrated in Figure 1c. The spectra of both activated carbons look similar, with the only difference observed in the spectrum of AC_CO2 at the peak at 1577.5 cm−1 assigned to C-N. This is in accordance with previous findings, as it has been reported that heteroatoms are removed during steam activation [30]. However, this functional group might impact the electrochemical behavior of the prepared carbons as it has been reported to improve wettability and conductivity with the carbon structure [31]. Besides, both activated carbons show peaks for O-H (3658 cm−1) and C-H groups like methylene or methyl (2982, 2890 and 1381 cm−1) as well as carboxylic groups (1253 cm−1), which have been reported for activated carbons in previous studies [32,33]. The double peak around 2349 cm−1, visible for both carbon samples, can be attributed to anti-symmetric stretching of atmospheric CO2 [34].
Figure 1d shows the Raman spectra of the activated carbons. Raman spectroscopy was used to analyze the structure, defects, and extent of graphitization in the porous activated carbon. Raman shifts were recorded for two prominent peaks around 1358 cm−1 for the D band and 1590 cm−1 for the G band. The D band is associated with the structural defects and disorder of the synthesized activated carbon, while the 2D mode corresponds to the second order of the D band [35]. The G band is also associated with the crystalline or graphitic nature of the as-prepared samples arising from the E2g vibrational mode and the in-plane stretching of the sp2-hybridized carbon atoms [36]. The D band is further deconvoluted and fitted into three peaks at 1354, 1522, and 1204 cm−1 (Figure S1a), attributed to D1, D3, and D4, respectively [37]. The intensity ratio of the D1 and G bands (ID1/IG) is calculated and used to assess the degree of structural disorderliness of the carbon samples [38]. The ID1/IG values for AC_CO2 and AC_H2O, calculated from the ratio of the peak heights, are 1.14 and 1.36, respectively. This is in accordance with the overall lower degree of activation and lower porosity of the AC_CO2 sample. A comparative analysis of the peak heights of the deconvoluted peaks according to the Lorentz function [39] is also given in Table 1 and Figure S1b.
The morphologies of carbonized and activated samples are shown in Figure 2. The surface of the precursor sample, SNS, as shown in Figure 2a, is relatively smooth and almost free of pores. As the precursor was carbonized in a nitrogen atmosphere at 900 °C, the surface of the sample was etched, which introduced voids to make the surface rough and irregular, as shown in Figure 2b. During the activation with CO2 and steam, as shown in Figure 2c,d, respectively, porous structures were developed for both samples. However, the AC_H2O-activated sample shows a more distorted surface structure, which can be attributed to the high activation energy of steam, which makes it easier to etch and penetrate the carbon structure whilst expanding the pores effectively.
Additionally, the structural characterization of the activated carbon samples was analyzed using nitrogen adsorption isotherms and recorded at 77 K (Figure 3). The calculated specific surface area (SBET), the average pore size, and total volume are reported in Table 2. Based on the IUPAC adsorption classifications, AC_CO2 exhibits a Type I isothermal characteristic of a microporous material with strong adsorbent–adsorptive interactions, whereas AC_H2O exhibits a Type II isotherm which indicates significant multilayer adsorption [40], as shown in Figure 3a. The sharp rise in adsorption within the low-pressure range of the adsorbed volume indicates the high-density micropores within the activated carbon structure. In contrast, AC_H2O exhibited higher adsorption than AC_CO2 in the low-pressure range. This is evident in the honeycomb structure, which has numerous mesopores and micropores developed by H2O activation. In Table 2, AC_H2O had a SSA of 1842 m2 g−1 and a total pore volume of 1.57 cm3 g−1 higher than 1528 m2 g−1 and 0.72 cm3 g−1 obtained by the AC_CO2. Again, this is attributed to steam’s diffusion effect in etching the pores’ internal layers to form numerous mesopores [28]. The pore size distribution of the activated carbons calculated using the non-localized density functional theory (NLDFT) and assuming slit-pores, presented in Figure 3b, shows that the sample AC_H2O has a significant share of pores in the mesopore range, while the porosity of AC_CO2 is almost completely limited to the micropore range below 2 nm. However, it has to be noted that due to the slower activation process and the higher yield in the CO2 process, there is potential for improving the porosity of the AC_CO2 sample by increasing the activation time or temperature.
XPS was further used to analyze the surface elemental components of the as-prepared carbon samples. The survey plot of AC_CO2 and AC_H2O is illustrated in Figure 4a, where two characteristic peaks for C1s and O1s were observed to be present in both samples. However, characteristic peaks for N1s and Ca2p were observed in the high-resolution spectra of AC_CO2 (Figure 4d). From Figure 4b, C is mostly bonded via single bonds to C or H (284.97 eV) and also bonded to O to form O=C-O (289.4 eV) [41]. O is also mainly bonded to carbon to form carboxyl bonds in the structure of the AC_CO2 sample at 531.93 eV, as seen in Figure 4c [42]. The remaining O atoms are bonded to other elements (Ca). Two unique peaks for N1s were also detected at 398.57 eV and 400.74 eV in the structure of AC_CO2, which correspond to pyridinic-N and pyrrolic-N, respectively (Figure 4d) [43]. These characteristic peaks enhance electron transfer at higher current densities [44]. The high-resolution spectra of C1s and O1s for AC_H2O are also presented in Figure 4e and Figure 4f, respectively. The intensity of the O1s peak shows that there is less O in the carbon structure of AC_H2O than AC_CO2, as evident from Table 3. This is because all the oxygen atoms bond to the carbon. The absence of N-functional groups in the AC_H2O sample can be attributed to the high energetic nature of the activation agent which removed most of the functional groups during activation, which was also reported by Breitenbach et al. [26]. The high-resolution spectra of Ca2p present in AC_CO2 is shown in Figure S2. The calculated elemental distributions are further provided in Table 3. The high amount of O and N in the AC_CO2 suggests that the sample is expected to show good retention capacity and conductivity in contrast to the AC_H2O sample despite having a larger surface area [43].

2.2. Electrochemical Characterization

The as-synthesized electrodes were tested in a symmetric two-electrode set-up using a Swagelok cell and 1M TEMA-BF4 in propylene carbonate as an electrolyte and measured by CV, GCD, and EIS. Figure 5a,b show the CV profiles from 1 mV s−1 to 300 mV s−1 within the potential range of −2.7 V to 2.7 V. The curves of the electrodes have a quasi-rectangular shape, which is typical behavior for double-layer supercapacitors. This shape depicts the fast-transport nature of the electrolyte ions and the energy storage capacity of the activated carbons. The curves obtained by the AC_H2O electrodes have a higher integral area than AC_CO2. As the scan rates increased, AC_CO2 maintained its quasi-rectangular shape through to 200 mV s−1 compared to AC_H2O, which only maintained its shape to the 100 mV s−1 scan rate. This is due to good reduction/oxidation reaction and fast ion responsiveness at higher scan rates by AC_CO2 electrodes, highlighting excellent charge storage and transport capabilities [13]. Figure 5c depicts the CV plot at a 10 mV s−1 scan rate as a comparative analysis for the as-prepared electrodes. It is evident from the plot that AC_H2O has a higher area than AC_CO2. Redox peaks are also observed at low scan rates and are more severe in AC_CO2. This might result from the residual N-containing compounds originating from the precursor (see Figure S3), which are present in the AC_CO2 sample but are washed out during water activation of the AC_H2O sample [30]. Due to the low N content (0.8 at. %), the effect of pseudo-capacitance is negligible, and the total capacitance values are dominated by the porosity, which is significantly higher for the AC_CO2 sample. Despite the absence of the N heteroatoms within the carbon structure of AC_H2O, it also displays humps in the CV profile. These can be attributed to the creation of a solid–electrolyte interphase (SEI) due to oxidation at high potentials that decreases the overall surface area of the activated carbon electrode and limits ion transport within the porous network, which ultimately reduces the specific capacitance with increased scan rates, as seen in Figure 5d [26].
The specific capacitance for the various scan rates is also reported in Figure 5d. The graph shows that the specific capacitance decreases with an increasing scan rate, which can be attributed to the insufficient time for ion diffusion at higher scan rates. AC_H2O achieves higher capacitance values from lower scan rates and drops sharply after 100 mV s−1. This is because, at higher potentials, electrolyte degradation tends to produce an SEI that reduces the overall surface area and consequently reduces the specific capacitance. In contrast, AC_CO2 shows good capacitance retentivity as it maintains its quasi-rectangular shape (Figure 5a) and capacitance at higher scan rates (Figure 5d). This phenomenon can be attributed to the presence of O and N heteroatoms in the AC_CO2 carbon structure, which tends to improve the conductivity and wettability of the electrode. The capacitance contributions for the devices were analyzed using Dunn’s method, considering the peak currents from various scan rates [45], and are reported in Figure 5e,f. Surface-controlled processes are more stable with increasing scan rates. In the AC_CO2 device, the capacitance is highly controlled by this mechanism and increases significantly as the scan rate increases. In contrast, diffusion-controlled processes are severe within the AC_H2O device. The pore structure of AC_H2O facilitates ion diffusion, and as such, at low scan rates, diffusion-controlled processes are more pronounced compared to the exact mechanism in the AC_CO2 electrode.
The capacitive nature of the synthesized electrodes was also investigated using the GCD technique at a current density from 0.1 to 4 A g−1 with a potential window of 2.7 V to −2.7 V, as shown in Figure 6a,b.
The GCD profiles for both electrodes exhibit nearly triangular charge–discharge curves at lower current densities, indicating ideal capacitive behavior. Again, the linearity of the discharge curve distorts at higher current densities, which can be attributed to internal resistance and poor ion transport efficiency [46]. From Figure 6c, the curve for AC_H2O shows a higher voltage drop than AC_CO2 during discharge, which indicates lower power density capability due to high resistance to ion diffusion. The specific capacitance of the electrodes is also calculated from the GCD results, as shown in Figure 6d. The specific capacitance for both electrodes decreases with increasing current density, typically due to the limited ion diffusion at higher currents. The AC_H2O electrode shows superior capacitance at lower current densities than the AC_CO2 electrode due to a more developed porous structure coupled with a high SSA. In contrast, AC_CO2 maintains its capacitance significantly. This characteristic behavior shows the excellent electrochemical property of AC_CO2 due to the presence of oxygenated and nitrogen functional groups.
Electrochemical impedance spectroscopy (EIS) measurement from 1 MHz to 10 mHz using a symmetric cell was performed to better understand the capacitive behavior and the ion transport within the electrode’s pores. The Nyquist plot in Figure 7a clearly shows distinct electrochemical characteristics: a steep line in the low-frequency region is indicative of their capacitive nature and the diameter of semi-circular arcs in the high-frequency region represents charge transfer resistance (RCT). In contrast, the intercept on the real axis represents the equivalent series resistance (Rs) [47]. The measured data was fitted to an equivalent circuit (inset of Figure 7a). The constant phase element, CPE1, is used to compensate for inhomogeneities in the system between electrode–electrolyte–current collector interphases. CPE2 is also employed to compensate for any pseudo-capacitance, while the Warburg impedance, Wo, also represents the total impedance due to the diffusion processes within the symmetric cell [48]. The Rs and Rct of AC_CO2 were calculated to be 4.65 Ω and 9.05 Ω, respectively, by fitting the circuit model shown in Figure 7a to a circuit. These values were lower than AC_H2O (6.87 Ω and 47.09 Ω for Rs and Rct, respectively). The Bode plot in Figure 7b also highlights the relationship between the impedance and frequency of porous activated carbon at high frequencies. Closer frequency response to a phase angle of −90° shows nearness to ideal capacitive behavior [49]. From Figure 7b, AC_CO2 shows a phase angle of −72.15° due to its improved conductive nature and wettability. In contrast, a lower phase angle (−56.48°) was recorded for AC_H2O due to the highly resistive nature of the electrode used for charge transfer. The capacitor frequency response at a phase angle of −45° shows the time constant (τ), which depicts the fast discharge nature of the supercapacitor device [50]. AC_CO2 has a low discharge time of 2.10 s compared to 5.73 s for AC_H2O.
To intrusively ascertain the practical application of the AC_CO2 symmetric device (Figure 7c), its cyclic stability is analyzed at 4 A g−1. The device maintained a good capacitance retention of 90% after 10,000 cycles, as shown in Figure 7d. A steady drop in retention was observed up to 7000 cycles; then, capacitance reached a plateau at approximately 90% of the initial value up to the end of the stability test at 10,000 cycles. The Ragone plot (Figure 7e) of AC_CO2 and AC_H2O symmetric devices displays a maximum energy density of 14.9 and 29.1 Wh kg−1, respectively. At a current density of 0.5 A g−1, energy densities of 9.5 Wh kg−1 and 9.2 Wh kg−1 at power densities of 299 W kg−1 and 204 W kg−1 were found for AC_CO2 and AC_H2O, respectively. These results are compared with other reported results at similar current densities of 0.5 A g−1 and 1 A g−1 (Figure 7f) [8,51,52,53,54,55,56].

3. Materials and Methods

3.1. Materials

SNSs, lignocellulosic biomass, were sourced from Bolgatanga in the Upper East region of Ghana. These shells served as the precursor material in the synthesis process. Graphite (99.5% pure, C-NERGY™, Imerys Graphite & Carbon Switzerland SA, Bodio, Switzerland), carbon-coated aluminum foil (z-flo 2651, Coveris Management GmbH, Vienna, Austria), polypropylene separator foil (Celgard® 3401, Celgard LLC, Charlotte, NC, USA), triethylmethylammonium tetrafluoroborate (98% pure, from TCI Deutschland GmbH, Eschborn, Germany), polytetrafluoroethylene (Fluorogistx CT LLC, Greenville, DE, USA), and propylene carbonate (99.5% pure, Acros Organics N.V., Geel, Belgium) were used for this experiment without modification. Distilled water (DW) was utilized throughout this work.

3.2. Activated Carbon Preparation

Figure 8 shows the schematic process for synthesizing activated carbon, similar to the work performed by Breitenbach et al. [26]. SNSs were washed with DW to remove any debris and dried in the sun for a day. The shells were ground using a universal mill (PULVERISETTE 19, FRITSCH GmbH, Idar-Oberstein, Germany) and sieved under 500 μm. Next, 50 g of the ground SNSs was dried in a vacuum overnight at 120 °C in an oven (Binder Inc., Bohemia, NY, USA), pyrolyzed in a nitrogen-rich chamber furnace (HTK8 Carbolite Gero GmbH, Neuhausen, Germany) using a heating rate of 5 K min−1, and held isothermally at a temperature of 900 °C for 30 min and left to cool to room temperature. Then, 10 g of carbonized SNSs was activated in a rotary furnace (RSR-B 120/500/11, Nabertherm GmbH, Lilienthal, Germany) at a temperature of 880 °C under a nitrogen atmosphere and kept isothermally for 30 min before actual activation with CO2 at a gas flow of 50 L h−1 for 120 min. Steam activation was also carried out using the same setup and program with a peristaltic pump attached using water at a flow rate of 1 mL min−1. The rotary tube was jacked up at one end of the rotary furnace with an installed cold trap to prevent condensed water from destroying the activated carbon. The samples activated with steam and CO2 are hereafter labeled as AC_H2O and AC_CO2, respectively. The yield of each thermal process was determined according to Equation (1).
Y = ω 0 ω a × 100 %
where w0 = the mass of carbon before carbonization or activation, wa = mass of carbon after carbonization or activation, and Y = obtained percentage yield of carbon.

3.3. Characterization of Lignocellulosic Biomass and Activated Carbon

The moisture content of the ground SNSs was measured using a moisture analyzer (MX-50, A&D Company, Tokyo, Japan). Thermal analysis of the precursor was performed with a STA449F3 TGA analyzer (NETZSCH, Selb, Germany). Carbonization, activation, and total yield were determined using Equation (1). The prepared samples’ specific surface area and total pore volume were investigated by gas adsorption using an Autosorb-iQ system (Anton Paar QuantaTec Inc., Graz, Austria). The morphology of the activated carbon and electrodes were characterized using a Phenom ProX scanning electron microscope (SEM) (Thermo Fisher Scientific, Waltham, MA, USA). Fourier transform infrared (FTIR) measurements were performed using a VERTEX70 spectrometer (PIKE GladiATR, Bruker Corporation, Billerica, MA, USA) in transmittance mode. Raman spectroscopy was also employed to examine the crystalline properties of the as-synthesized carbon samples using a HORIBA XploRA (HORIBA France SAS, Palaiseau, France) with a 532 mm excitation wavelength and a grating of 1800 grooves/mm. The surface functional groups were further investigated via X-ray photoelectron spectroscopy using a NEXSA G2 XPS system (Thermo Fisher Scientific, Waltham, MA, USA). The samples were probed by an Al-Kα monochromatic X-ray source (1486.6 eV) using a spot size of 400 μm in diameter. The hemispherical analyzer was operated in a constant analyzer energy mode with a pass energy of 200 eV and energy step of 1 eV for the survey spectra, whereas 20 eV and 0.05 eV steps were taken for recording the high-resolution spectra. A dual flood gun, which provides low-energy electrons and Ar ions as a simultaneous beam, was used to compensate the surface charge. A standard charge shift referencing the spectra via a C1s peak of adventitious carbon at 285.0 eV was applied.

3.4. Electrode Fabrication and Characterization

The working electrode was prepared by mixing as-synthesized activated carbon, graphite, and PTFE in an 80:10:10 mass ratio into a kneadable dough of 100 µm thickness. To ensure optimal performance, 8 mm electrodes were punched out from the dough and dried in a vacuum overnight at 110 °C. The electrochemical performance of the fabricated electrodes was evaluated using a symmetric two-electrode system. For this setup, two electrodes of similar masses, a 9 mm separator, two 8 mm carbon-coated aluminum foils serving as current collectors, and 200 µL of 1 M TEMA-BF4 solution in propylene carbonate as the electrolyte were assembled into Swagelok®-type test cells (Figure S4). The schematic for the cell assembly is provided in Figure S5. Electrochemical impedance spectroscopy (EIS), cyclic voltammetry (CV), and galvanostatic charge–discharge (GCD) analyses were performed using a multi-channel electrochemical potentiostat (Vertex. One, IviumTechnologies BV, Eindhoven, The Netherlands).
EIS measurements were carried out with an AC voltage amplitude of 10 mV, spanning a frequency range of 1 MHz to 0.01 Hz. Specific capacitance (Cs), energy density (Es), and power density (Ps) were calculated using data from CV and GCD experiments. These calculations considered the total mass (twice the mass of active material) and the single mass of the active material (80% of the electrode’s mass) for the device and electrode analysis, respectively. The equations for the analysis are referenced below:
C s , C v = V i V f i = d V 2 m E · r
C s , G C D = 2 · I · t m E Δ V
C s , C v   in Equation (2) is the specific capacitance of the electrode, Vf and Vi are the initial and final voltages of the potential window ΔV, r is the scan rate, and mE is the active mass of the electrode as reported by Breitenbach and co-researchers in 2021 [26]. Equation (3) was used to calculate the specific capacitance from the GCD results, where the current = I, active mass of electrode = mE, discharge time = t, and potential window = ΔV. Equations (4) and (5) were also used to calculate energy density, Es, and power density, Ps, for the symmetric device.
E s = 1 8 · C s , G D C ( Δ V )   2
P s = E s t

4. Conclusions

This study successfully demonstrates the potential of SNSs as a sustainable source for producing activated carbon electrodes in a two-stage activation process for supercapacitor applications. CO2 and H2O steam were used as physical activation agents for synthesis. The analysis showed that steam activation created a highly porous structure with a larger SSA (1842 m2 g−1) and higher total pore volume (1.57 cm3 g−1). These properties of AC_H2O translated into superior specific capacitance of 102.2 F g−1 at 0.5 A g−1, coupled with an energy density of 9.2 Wh kg−1 at a higher power density of 204 W kg−1 in a two-electrode symmetric Swagelok cell. In contrast, CO2 activation produced carbon with a higher yield but reasonable SSA of 1528 m2 g−1, thus showing potential for further improvements through optimized activation time and/or temperature. The sample already shows better graphitization degree, a lower charge transfer resistance of 9.7 Ω, and the presence of O and N heteroatom functional groups. The AC_CO2 symmetric device delivers a specific capacitance of 47.5 F g−1 at 0.5 A g−1 and an energy density of 9.5 Wh kg−1 at 199 W kg−1 power density, and it also achieved 90% capacitance retention after 10,000 cycles. This research highlights an innovative way to repurpose agricultural waste into high-performance energy storage materials, contributing to sustainable waste management and renewable energy solutions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics13070224/s1, Figure S1: High resolution XPS spectra of Ca2p in AC_CO2; Figure S2: Schematic of the Swagelok cell; Figure S3: Schematic of the supercapacitor cell preparation. Figure S4: Schematic of the Swagelok cell. Figure S5: Schematic of the supercapacitor cell preparation.

Author Contributions

J.A.A.: conceptualization, methodology, investigation, data curation, visualization, and writing—original draft. N.N.: methodology and investigation. D.N.A.: formal analysis and writing—review and editing. F.O.A.: visualization and writing—review and editing. M.A.-A.: data curation and writing—review and editing. R.O.: validation and writing—review and editing. R.Z.: investigation and writing—review and editing. C.U.: conceptualization, methodology, validation, resources, supervision, data curation, funding acquisition, and writing—review and editing. K.M.-D.: conceptualization, funding acquisition, validation, supervision, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

The authors thank the Kwame Nkrumah University of Science and Technology Engineering Education Project (KEEP) for the financial support. Further financial support from the European Regional Development Fund (EFRE) and the province of Upper Austria through the program IBW 2021–2027 (Project Sus2C) is gratefully acknowledged.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

Authors Joshua Atta Alabi, Neda Nazari and Christoph Unterweger were employed by the company Wood K plus—Kompetenzzentrum Holz GmbH. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Singh, S.; Singh, S. Advancements and challenges in integrating renewable energy sources into distribution grid systems: A comprehensive review. J. Energy Resour. Technol. 2024, 146, 090801. [Google Scholar] [CrossRef]
  2. Khalid, M. Smart grids and renewable energy systems: Perspectives and grid integration challenges. Energy Strategy Rev. 2024, 51, 101299. [Google Scholar]
  3. Oyedotun, K.O.; Mamba, B.B. New trends in supercapacitors applications. Inorg. Chem. Commun. 2024, 170, 113154. [Google Scholar] [CrossRef]
  4. Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [PubMed]
  5. Girirajan, M.; Bojarajan, A.K.; Pulidindi, I.N.; Hui, K.N.; Sangaraju, S. An insight into the nanoarchitecture of electrode materials on the performance of supercapacitors. Coord. Chem. Rev. 2024, 518, 216080. [Google Scholar] [CrossRef]
  6. Ampong, D.N.; Effah, E.; Tsiwah, E.A.; Kumar, A.; Agyekum, E.; Doku, E.N.A.; Issaka, O.; Agyemang, F.O.; Mensah-Darkwa, K.; Gupta, R.K. Advances and challenges in covalent organic frameworks as an emerging class of materials for energy and environmental concerns. Coord. Chem. Rev. 2024, 519, 216121. [Google Scholar]
  7. Ampong, D.N.; Mensah-Darkwa, K.; Gupta, R.K. Bio-Based Carbon for Supercapacitors, in NanoCarbon: A Wonder Material for Energy Applications: Volume 2: Fundamentals and Advancement for Energy Storage Applications; Springer: Berlin/Heidelberg, Germany, 2024; pp. 261–279. [Google Scholar]
  8. Ampong, D.N.; Lin, W.; de Souza, F.M.; Bharti, V.K.; Agyemang, F.O.; Andrews, A.; Mensah-Darkwa, K.; Dhakal, A.; Mishra, S.R.; Perez, F.; et al. Utilization of shea butter waste-derived hierarchical activated carbon for high-performance supercapacitor applications. Bioresour. Technol. 2024, 406, 131039. [Google Scholar]
  9. Devi, M.S.; Thangadurai, T.D.; Shanmugaraju, S.; Selvan, C.P.; Lee, Y.I. Biomass waste from walnut shell for pollutants removal and energy storage: A review on waste to wealth transformation. Adsorption 2024, 30, 891–913. [Google Scholar] [CrossRef]
  10. Singla, M.K.; Gupta, J.; Safaraliev, M.; Nijhawan, P.; Oberoi, A.S.; Menaem, A.A. Characterization of an activated carbon electrode made from coconut shell precursor for hydrogen storage applications. Int. J. Hydrogen Energy 2024, 61, 1417–1428. [Google Scholar] [CrossRef]
  11. Le Van, K.; Thi, T.T.L. Activated carbon derived from rice husk by NaOH activation and its application in supercapacitor. Prog. Nat. Sci. Mater. Int. 2014, 24, 191–198. [Google Scholar] [CrossRef]
  12. Abdallah, F.; Arthur, E.K.; Mensah-Darkwa, K.; Gikunoo, E.; Baffour, S.A.; Agamah, B.A.; Nartey, M.A.; Agyemang, F.O. Electrochemical performance of corncob-derived activated carbon-graphene oxide and TiO2 ternary composite electrode for supercapacitor applications. J. Energy Storage 2023, 68, 107776. [Google Scholar] [CrossRef]
  13. Zhan, Y.; Zhou, H.; Guo, F.; Tian, B.; Du, S.; Dong, Y.; Qian, L. Preparation of highly porous activated carbons from peanut shells as low-cost electrode materials for supercapacitors. J. Energy Storage 2021, 34, 102180. [Google Scholar] [CrossRef]
  14. Jain, A.; Tripathi, S. Nano-porous activated carbon from sugarcane waste for supercapacitor application. J. Energy Storage 2015, 4, 121–127. [Google Scholar] [CrossRef]
  15. Shen, H.-H.; Hu, C.-C. A high-voltage asymmetric electrical double-layer capacitors using propylene carbonate. Electrochem. Commun. 2016, 70, 23–27. [Google Scholar] [CrossRef]
  16. Nanzumani, N.M.; Agyemang, F.O.; Mensah-Darkwa, K.; Appiah, E.S.; Arthur, E.K.; Gikunoo, E.; Koomson, B.; Jadhav, A.R.; Raji, A. Molten salt synthesis of nitrogen-doped hierarchical porous carbon from plantain peels for high performance supercapacitor. J. Electroanal. Chem. 2022, 920, 116645. [Google Scholar] [CrossRef]
  17. Misnon, I.I.; Zain, N.K.M.; Jose, R. Conversion of oil palm kernel shell biomass to activated carbon for supercapacitor electrode application. Waste Biomass Valorization 2019, 10, 1731–1740. [Google Scholar] [CrossRef]
  18. Taer, E.; Apriwandi, A.; Ningsih, Y.; Taslim, R. Preparation of activated carbon electrode from pineapple crown waste for supercapacitor application. Int. J. Electrochem. Sci. 2019, 14, 2462–2475. [Google Scholar] [CrossRef]
  19. Ospino, J.; Parra-Barraza, J.P.; Cervera, S.; Coral-Escobar, E.E.; Vargas-Ceballos, O.A. Activated carbon from cassava peel: A promising electrode material for supercapacitors. Rev. Fac. De Ing. Univ. De Antioq. 2022, 102, 88–95. [Google Scholar]
  20. Kumar, R.; Singh, K.; Kumar, P.; Kaur, A. Highly porous activated carbon prepared from the bio-waste of yellow mustard seed for high-capacity supercapacitor applications. Mater. Chem. Phys. 2023, 304, 127869. [Google Scholar] [CrossRef]
  21. Subramani, K.; Sudhan, N.; Karnan, M.; Sathish, M. Orange peel derived activated carbon for fabrication of high-energy and high-rate supercapacitors. ChemistrySelect 2017, 2, 11384–11392. [Google Scholar] [CrossRef]
  22. Adazabra, A.N.; Viruthagiri, G.; Shanmugam, N. Management of spent shea waste: An instrumental characterization and valorization in clay bricks construction. Waste Manag. 2017, 64, 286–304. [Google Scholar] [PubMed]
  23. Agwuncha, S.C.; Owonubi, S.; Fapojuwo, D.P.; Abdulkarim, A.; Okonkwo, T.P.; Makhatha, E.M. Evaluation of mercerization treatment conditions on extracted cellulose from shea nut shell using FTIR and thermogravimetric analysis. Mater. Today Proc. 2020, 38, 958–963. [Google Scholar] [CrossRef]
  24. Abdullahi, Z.; Ajai, A.; Iyaka, Y. Removal of Lead and Cadmium from Aqueous Solution Using Activated Carbons from Desert Dates and Shea Nut Shells. 2017. Available online: http://irepo.futminna.edu.ng:8080/jspui/bitstream/123456789/9342/1/22Article%206.pdf (accessed on 17 June 2025).
  25. Musah, M.; Matthew, J.T.; Azeh, Y.; Badeggi, U.M.; Muhammad, A.I.; Abu, L.M.; Okonkwo, P.T.; Muhammad, K.T. Preparation and characterization of adsorbent from waste shea (vitellaria paradoxa) nut shell. Fudma J. Sci. 2024, 8, 338–344. [Google Scholar] [CrossRef]
  26. Breitenbach, S.; Lumetzberger, A.; Hobisch, M.A.; Unterweger, C.; Spirk, S.; Stifter, D.; Fürst, C.; Hassel, A.W. Supercapacitor Electrodes from Viscose-Based Activated Carbon Fibers: Significant Yield and Performance Improvement Using Diammonium Hydrogen Phosphate as Impregnating Agent. C 2020, 6, 17. [Google Scholar] [CrossRef]
  27. Zou, Y.; Wang, H.; Xu, L.; Dong, M.; Shen, B.; Wang, X.; Yang, J. Synergistic effect of CO2 and H2O co-activation of Zhundong coal at a low burn-off rate on high performance supercapacitor. J. Power Sources 2023, 556, 232509. [Google Scholar] [CrossRef]
  28. Li, L.; Tong, S.; Duan, L.; Zhao, C.; Shi, Z. Effect of CO2 and H2O on lignite char structure and reactivity in a fluidized bed reactor. Fuel Process. Technol. 2021, 211, 106564. [Google Scholar] [CrossRef]
  29. Wei, R.; Ren, L.; Geng, F. Gasification reactivity and characteristics of coal chars and petcokes. J. Energy Inst. 2021, 96, 25–30. [Google Scholar] [CrossRef]
  30. Breitenbach, S.; Duchoslav, J.; Mardare, A.I.; Unterweger, C.; Stifter, D.; Hassel, A.W.; Fürst, C. Comparative Behavior of Viscose-Based Supercapacitor Electrodes Activated by KOH, H2O, and CO2. Nanomaterials 2022, 12, 677. [Google Scholar] [CrossRef]
  31. Yang, H.; Yan, R.; Chen, H.; Lee, D.H.; Liang, D.T.; Zheng, C. Mechanism of palm oil waste pyrolysis in a packed bed. Energy Fuels 2006, 20, 1321–1328. [Google Scholar] [CrossRef]
  32. Yang, L.; Jiang, X.; Huang, T.; Jiang, W. Physicochemical characteristics and desulphurization activity of pyrolusite-blended activated coke. Environ. Technol. 2015, 36, 2847–2854. [Google Scholar] [CrossRef]
  33. Zhang, Y.J.; Xing, Z.J.; Duan, Z.K.; Li, M.; Wang, Y. Effects of steam activation on the pore structure and surface chemistry of activated carbon derived from bamboo waste. Appl. Surf. Sci. 2014, 315, 279–286. [Google Scholar] [CrossRef]
  34. Schott, J.A.; Do-Thanh, C.-L.; Shan, W.; Puskar, N.G.; Dai, S.; Mahurin, S.M. FTIR investigation of the interfacial properties and mechanisms of CO2 sorption in porous ionic liquids. Green Chem. Eng. 2021, 2, 392–401. [Google Scholar] [CrossRef]
  35. Le Ru, E.C.; Auguié, B. Enhancement factors: A central concept during 50 years of surface-enhanced Raman spectroscopy. ACS Nano 2024, 18, 9773–9783. [Google Scholar] [CrossRef] [PubMed]
  36. Cançado, L.G.; Monken, V.P.; Campos, J.L.E.; Santos, J.C.; Backes, C.; Chacham, H.; Neves, B.R.A.; Jorio, A. Science and metrology of defects in graphene using Raman spectroscopy. Carbon 2024, 220, 118801. [Google Scholar] [CrossRef]
  37. Wu, D.; Sun, F.; Qu, Z.; Wang, H.; Lou, Z.; Wu, B.; Zhao, G. Multi-scale structure optimization of boron-doped hard carbon nanospheres boosting the plateau capacity for high performance sodium ion batteries. J. Mater. Chem. A 2022, 10, 17225–17236. [Google Scholar] [CrossRef]
  38. Nazari, N.; Abadi, M.D.M.; Khachatourian, A.M.; Golmohammad, M.; Nemati, A. The effect of phosphorus and nitrogen dopants on structural, microstructural, and electrochemical characteristics of 3D reduced graphene oxide as an efficient supercapacitor electrode material. Diam. Relat. Mater. 2023, 137, 110144. [Google Scholar] [CrossRef]
  39. Zhang, X.; Hou, Z.; Jiang, M.; Peng, J.; Ma, H.; Gao, Y.; Wang, J.G. Molecular Engineering to Regulate the Pseudo-Graphitic Structure of Hard Carbon for Superior Sodium Energy Storage. Small 2024, 20, 2311778. [Google Scholar] [CrossRef]
  40. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef]
  41. Tan, Z.; Li, H.; Feng, Q.; Jiang, L.; Pan, H.; Huang, Z.; Shuai, M.; Kuang, Y. One-pot synthesis of Fe/N/S-doped porous carbon nanotubes for efficient oxygen reduction reaction. J. Mater. Chem. A 2019, 7, 1607–1615. [Google Scholar] [CrossRef]
  42. Du, J.; Chen, A.; Yu, Y.; Zhang, Y.; Lv, H.; Liu, L. Mesoporous carbon sheets embedded with vesicles for enhanced supercapacitor performance. J. Mater. Chem. A 2019, 7, 15707–15713. [Google Scholar] [CrossRef]
  43. Wang, M.; Liu, B.; Chen, H.; Yang, D.; Li, H. N/O codoped porous carbons with layered structure for high-rate performance supercapacitors. ACS Sustain. Chem. Eng. 2019, 7, 11219–11227. [Google Scholar] [CrossRef]
  44. Lei, W.; Guo, J.; Wu, Z.; Xuan, C.; Xiao, W.; Wang, D. Highly nitrogen and sulfur dual-doped carbon microspheres for supercapacitors. Sci. Bull. 2017, 62, 1011–1017. [Google Scholar] [CrossRef]
  45. Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive oxide materials for high-rate electrochemical energy storage. Energy Environ. Sci. 2014, 7, 1597–1614. [Google Scholar] [CrossRef]
  46. Ike, I.S.; Sigalas, I.; Iyuke, S. Understanding performance limitation and suppression of leakage current or self-discharge in electrochemical capacitors: A review. Phys. Chem. Chem. Phys. 2016, 18, 661–680. [Google Scholar] [CrossRef]
  47. Qi, F.; Xia, Z.; Sun, R.; Sun, X.; Xu, X.; Wei, W.; Wang, S.; Sun, G. Graphitization induced by KOH etching for the fabrication of hierarchical porous graphitic carbon sheets for high performance supercapacitors. J. Mater. Chem. A 2018, 6, 14170–14177. [Google Scholar] [CrossRef]
  48. Pina, A.C.; Amaya, A.; Marcuzzo, J.S.; Rodrigues, A.C.; Baldan, M.R.; Tancredi, N.; Cuña, A. Supercapacitor electrode based on activated carbon wool felt. C 2018, 4, 24. [Google Scholar] [CrossRef]
  49. Krishnamoorthy, K.; Veerasubramani, G.K.; Radhakrishnan, S.; Kim, S.J. Supercapacitive properties of hydrothermally synthesized sphere like MoS2 nanostructures. Mater. Res. Bull. 2014, 50, 499–502. [Google Scholar] [CrossRef]
  50. Liu, Z.; Xiao, K.; Guo, H.; Ning, X.; Hu, A.; Tang, Q.; Fan, B.; Zhu, Y.; Chen, X. Nitrogen-doped worm-like graphitized hierarchical porous carbon designed for enhancing area-normalized capacitance of electrical double layer supercapacitors. Carbon 2017, 117, 163–173. [Google Scholar] [CrossRef]
  51. Ma, G.; Li, J.; Sun, K.; Peng, H.; Feng, E.; Lei, Z. Tea-leaves based nitrogen-doped porous carbons for high-performance supercapacitors electrode. J. Solid State Electrochem. 2017, 21, 525–535. [Google Scholar] [CrossRef]
  52. Taurbekov, A.; Abdisattar, A.; Atamanov, M.; Yeleuov, M.; Daulbayev, C.; Askaruly, K.; Kaidar, B.; Mansurov, Z.; Castro-Gutierrez, J.; Celzard, A.; et al. Biomass Derived High Porous Carbon via CO2 Activation for Supercapacitor Electrodes. J. Compos. Sci. 2023, 7, 444. [Google Scholar] [CrossRef]
  53. Zuo, S.; Chen, J.; Liu, W.; Li, X.; Kong, Y.; Yao, C.; Fu, Y. Preparation of 3D interconnected hierarchical porous N-doped carbon nanotubes. Carbon 2018, 129, 199–206. [Google Scholar] [CrossRef]
  54. Yakaboylu, G.A.; Jiang, C.; Yumak, T.; Zondlo, J.W.; Wang, J.; Sabolsky, E.M. Engineered hierarchical porous carbons for supercapacitor applications through chemical pretreatment and activation of biomass precursors. Renew. Energy 2021, 163, 276–287. [Google Scholar] [CrossRef]
  55. Cao, Y.; Han, K.; Teng, Z.; Li, J.; Ji, T.; Li, X.; Zhang, J. Optimized synergistic preparation of nitrogen-doped porous carbon derived from gasified carbon for supercapacitors. J. Alloys Compd. 2021, 860, 158385. [Google Scholar] [CrossRef]
  56. Gao, M.; Wang, L.; Zhao, B.; Gu, X.; Li, T.; Huang, L.; Wu, Q.; Yu, S.; Liu, S. Sandwich construction of chitosan/reduced graphene oxide composite as additive-free electrode material for high-performance supercapacitors. Carbohydr. Polym. 2021, 255, 117397. [Google Scholar]
Figure 1. Physicochemical properties of activated carbon: (a) TGA and DTGA curve of precursor SNS, (b) carbon yield analysis plot, (c) FTIR plot, and (d) Raman spectra.
Figure 1. Physicochemical properties of activated carbon: (a) TGA and DTGA curve of precursor SNS, (b) carbon yield analysis plot, (c) FTIR plot, and (d) Raman spectra.
Inorganics 13 00224 g001
Figure 2. SEM micrograph of (a) SNS, (b) C_SNS, (c) AC_CO2, and (d) AC_H2O.
Figure 2. SEM micrograph of (a) SNS, (b) C_SNS, (c) AC_CO2, and (d) AC_H2O.
Inorganics 13 00224 g002
Figure 3. (a) N2 adsorption isotherms. (b) Pore size distribution plot of activated carbons.
Figure 3. (a) N2 adsorption isotherms. (b) Pore size distribution plot of activated carbons.
Inorganics 13 00224 g003
Figure 4. (a) XPS survey spectra of activated carbons: (b) C1s, (c) O1s, and (d) N1s spectra of AC_CO2; (e) C1s and (f) O1s XPS spectra of AC_H2O.
Figure 4. (a) XPS survey spectra of activated carbons: (b) C1s, (c) O1s, and (d) N1s spectra of AC_CO2; (e) C1s and (f) O1s XPS spectra of AC_H2O.
Inorganics 13 00224 g004
Figure 5. CV curves at various scan rates for (a) AC_CO2 and (b) AC_H2O, (c) at 10 mV s−1 for both electrodes, (d) specific capacitance vs. scan rate plot, and capacitance contribution of (e) AC_CO2 and (f) AC_H2O.
Figure 5. CV curves at various scan rates for (a) AC_CO2 and (b) AC_H2O, (c) at 10 mV s−1 for both electrodes, (d) specific capacitance vs. scan rate plot, and capacitance contribution of (e) AC_CO2 and (f) AC_H2O.
Inorganics 13 00224 g005
Figure 6. GCD plots at 0.1 A g−1–5 A g−1 for (a) AC_H2O and (b) AC_CO2; (c) GCD plot for both electrodes at 0.5 A g−1; (d) specific capacitance of electrodes at various current densities.
Figure 6. GCD plots at 0.1 A g−1–5 A g−1 for (a) AC_H2O and (b) AC_CO2; (c) GCD plot for both electrodes at 0.5 A g−1; (d) specific capacitance of electrodes at various current densities.
Inorganics 13 00224 g006
Figure 7. EIS: (a) Nyquist plot; (b) Bode plot of AC_CO2 and AC_H2O symmetric devices; (c) Swagelok cell assembly; (d) cycle stability plot; (e) Ragone plot of devices; (f) comparative analysis with literature [8,51,52,53,54,55,56].
Figure 7. EIS: (a) Nyquist plot; (b) Bode plot of AC_CO2 and AC_H2O symmetric devices; (c) Swagelok cell assembly; (d) cycle stability plot; (e) Ragone plot of devices; (f) comparative analysis with literature [8,51,52,53,54,55,56].
Inorganics 13 00224 g007
Figure 8. Schematic for the activated carbon preparation from SNS.
Figure 8. Schematic for the activated carbon preparation from SNS.
Inorganics 13 00224 g008
Table 1. Structural properties obtained from Raman analysis of physically activated carbons.
Table 1. Structural properties obtained from Raman analysis of physically activated carbons.
Sample IDID1 (a.u)ID3 (a.u)ID4 (a.u)IGID1/IG
AC_CO2159550615413951.14
AC_H2O227574043816691.36
Table 2. Textural properties of activated carbon samples.
Table 2. Textural properties of activated carbon samples.
SampleSBET (m2 g−1)VT (cm3 g−1)Pore Diameter (nm)
AC_CO215280.721.95
AC_H2O18421.573.41
Table 3. Elemental composition from high-resolution XPS.
Table 3. Elemental composition from high-resolution XPS.
Sample IDElemental Composition (at. %)
CONCa
AC_CO281.115.10.83.0
AC_H2O96.53.5--
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alabi, J.A.; Nazari, N.; Ampong, D.N.; Agyemang, F.O.; Adom-Asamoah, M.; Opoku, R.; Zahrhuber, R.; Unterweger, C.; Mensah-Darkwa, K. Synthesis of Physically Activated Carbons from Vitellaria paradoxa Shells for Supercapacitor Electrode Applications. Inorganics 2025, 13, 224. https://doi.org/10.3390/inorganics13070224

AMA Style

Alabi JA, Nazari N, Ampong DN, Agyemang FO, Adom-Asamoah M, Opoku R, Zahrhuber R, Unterweger C, Mensah-Darkwa K. Synthesis of Physically Activated Carbons from Vitellaria paradoxa Shells for Supercapacitor Electrode Applications. Inorganics. 2025; 13(7):224. https://doi.org/10.3390/inorganics13070224

Chicago/Turabian Style

Alabi, Joshua Atta, Neda Nazari, Daniel Nframah Ampong, Frank Ofori Agyemang, Mark Adom-Asamoah, Richard Opoku, Rene Zahrhuber, Christoph Unterweger, and Kwadwo Mensah-Darkwa. 2025. "Synthesis of Physically Activated Carbons from Vitellaria paradoxa Shells for Supercapacitor Electrode Applications" Inorganics 13, no. 7: 224. https://doi.org/10.3390/inorganics13070224

APA Style

Alabi, J. A., Nazari, N., Ampong, D. N., Agyemang, F. O., Adom-Asamoah, M., Opoku, R., Zahrhuber, R., Unterweger, C., & Mensah-Darkwa, K. (2025). Synthesis of Physically Activated Carbons from Vitellaria paradoxa Shells for Supercapacitor Electrode Applications. Inorganics, 13(7), 224. https://doi.org/10.3390/inorganics13070224

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