Next Article in Journal
Investigation of Materials for Enhanced Vibrational Insulation in Ground Structures Using Concrete Mixtures: A Case Study of Rubber Aggregate Addition
Previous Article in Journal
Nanocones: A Compressive Review of Their Electrochemical Synthesis and Applications
Previous Article in Special Issue
Lightweight 3D Lithiophilic Graphene Aerogel Current Collectors for Lithium Metal Anodes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Pecan Shell-Derived Activated Carbon for High-Electrochemical Performance Supercapacitor Electrode

1
Department of Electrical Engineering, Stanford University, Stanford, CA 94305, USA
2
Department of Materials Science and Engineering, University of North Texas, Denton, TX 76207, USA
3
Department of Mechanical Engineering, University of North Texas, Denton, TX 76207, USA
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Materials 2024, 17(13), 3091; https://doi.org/10.3390/ma17133091
Submission received: 25 April 2024 / Revised: 12 June 2024 / Accepted: 19 June 2024 / Published: 24 June 2024

Abstract

:
Carbon nanomaterials-based electric double-layer capacitors (EDLCs) are reliable and appealing energy-storage systems offering high power density and long cycling stability. However, these energy storage devices are plagued with critical shortcomings, such as low specific capacitance, inefficient physical/chemical activation process, and self-discharge of electrode materials, hindering their future application. In this work, we use a self-activation process, an environmentally benign and low-cost process, to produce high-performance activated carbon (AC). Novel activated carbon from pecan shells (PS) was successfully synthesized through a single-step self-activation process, which combines the carbonization and activation processes. The as-synthesized pecan shell-derived activated carbon (PSAC) provides a high-porosity, low-resistance, and ordered pore structure with a specific pore volume of 0.744 cm3/g and BET surface area of 1554 m2/g. The supercapacitors fabricated from PSAC demonstrate a specific capacitance of 269 F/g at 2 A/g, excellent cycling stability over 15,000 cycles, and energy and power density of 37.4 Wh/kg and of 2.1 kW/kg, respectively. It is believed that the high-efficiency PSAC synthesized from the novel self-activation method could provide a practical route to environmentally friendly and easily scalable supercapacitors.

1. Introduction

The increasing need for environmentally benign and sustainable energy solutions has spurred significant research into renewable energy storage systems [1]. Among the available energy-storage systems, supercapacitors, with their incredibly high power density, ultra-fast charging/discharging rate (<60 s), high cycle life (>100,000), and excellent stability compared to traditional capacitors and lithium-ion batteries, are an essential candidate to power future portable and flexible electronics, electric vehicles, and emergency door opening in airplanes [1,2,3]. To meet the demands of green energy capture and usage, supercapacitor performance needs to be improved [4]. Recent research has significantly improved the electrochemical performance of supercapacitors through the design and synthesis of novel carbon-based electrode materials [2,3]; however, these are often produced by non-renewable energy sources and involve environmentally hazardous processes. Therefore, further breakthroughs are needed to produce high-efficiency carbon materials by using renewable resources and an environmentally benign processes [5].
Over the past ten years, scientists have explored various carbon-based substances, including carbon nanotubes (CNTs), mesoporous carbon, graphene, and carbon nanofibers (CNFs), as potential materials for supercapacitors. This is attributed to their high specific surface area (SSA), electrical conductivity, and electrochemical stability. The existing commercial precursor for producing activated carbon originates from non-renewable energy sources. As an alternative, biomass-based ACs have been derived from nut shells including corn stalk core [6], coffee shell [7], hemp [8], human hair [9], and methylene blue [10]. Nutshells, which contains high amount of lignin and ordered cellular structure, have been widely utilized as a precursor for synthesizing large-scale activated carbon with a high product yield (>60%) [11,12]. Further, activated carbon from nutshells demonstrate excellent physical properties including high-surface area (>1000 m3/g) and an interconnected pore structure [6,7,8,9,10]. Similarly, pecan shells are an attractive candidate for biomass-activated carbon material because of its abundance—with 246 million pounds produced as agriculture waste per year in the United States. Since the amount of agricultural waste biomass being generated is predicted to increase with the continued growth of agricultural activities, it is important to generate useful products from these wastes. ACs from pecan shell have been investigated for use as a heavy metal ion water filter and carbon dioxide absorber [13,14,15]. However, the use of ACs from pecan shell as a supercapacitor electrode have not been reported. Further, few physical and chemical activation processes have been developed to synthesize activated carbon from nut shells; these methods involve toxic reagents such as KOH, ZnCl2, H3PO4, H2SO4, and strong acids that are later released into the environment [15,16,17,18,19,20,21]. Recently, a self-activation process was developed as an alternative to these toxic physical and chemical processes for fabricating high performance activated carbon by our research group [22,23,24,25]. Self-activation utilizes the in situ gases generated during the heat treatment to activate the carbon in biomass [22]; since no additional activating agent was used, this self-activation process is environmentally benign and economical.
Here, we report an efficient synthesis and self-activation of the activated carbon (AC) from pecan shells for the application of supercapacitor electrode. The as-synthesized pecan shell activated carbon (PSAC) exhibited a high specific pore volume with narrow pore distribution, and high specific BET surface area. These appealing characteristics of the PSAC result in a high specific capacitance, high power density, and excellent cycling stability of the supercapacitor. The outstanding electrochemical performance of these PSACs electrodes could help meet the growing demand for sustainable and scalable fabrication of supercapacitors.

2. Experimental

2.1. Activation of Pecan Shells

Locally grown pecans were collected and shelled. The pecan shells were placed in an open ceramic crucible and introduced into a versatile box furnace (STY-1600C, Sentro Tech Corp., Strongsville, OH, USA). Subsequently, the furnace door was sealed, and a vacuum was applied until a pressure of −730 torr (96% vacuum) was reached. No inert gases were used in the activation during the self-activation process. Then, the furnace’s air outtake valve was sealed, preventing gas exchange with the external environment. The temperature was then increased with a maximum ramping rate of 10 °C/min to 1050 °C. At 1050 °C, the sample was left to dwell for either 5 h, 10 h, and 15 h. Hereafter, we refer to the samples with the heat treatment at set dwelling times as PSAC-5, PSAC-10, and PSAC-15, respectively. During this dwelling time the samples in the furnace underwent self-activation. After dwelling for the specified amount of time, the furnace was then cooled to room temperature at a rate of 10 °C/min, where the internal environment of the furnace was naturally cooled to ambient temperature. Then, they were removed from the furnace and small samples were taken for surface characterization. The rest were ground into a powder for further processing.

2.2. Physical Characterization

We conducted nitrogen adsorption and desorption to measure the specific surface area, pore volume, and pore size distribution of the PSACs through the Brunauer–Emmett–Teller (BET) method (3-Flex 3500, Micromeritics Instrument Corp., Norcross, GA, USA) [24]. The PSAC sample was placed into a test tube and degassed at 350 °C in a degasser for one day (VacPrep 061, Micromeritics Instrument Corp.). Nitrogen adsorption and desorption experiments were conducted at various relative pressures over the course of one day to measure the surface area of the PSAC. The assessment of mesopores-macropore surface area and pore volume was performed using the Barrett–Joyner–Halenda (BJH) method (3-Flex 3500, Micromeritics Instrument Corp.). Additionally, micropore surface area was assessed by the Horvath–Kawazoe (HK) method, and micropore pore volume was calculated by t-plot. Raman spectroscopy (Nicolet Almega XR Dispersive Raman spectrometer, Thermo Scientific, Waltham, MA, USA)) was used to characterize the carbon peak on the compressed powder PSAC samples at a range of 1000–3000 cm−1. Morphologies of the as-fabricated PSAC were measured and assessed by using scanning electron microscope (FEI Nova-NanoSEM 230, Field Electron and Ion Company, Hillsboro, OR, USA). Microstructure investigation was performed through transmission electron microscopy (Philips EM420 TEM, Philips, Amsterdam, The Netherlands).

2.3. Electrochemical Characterization

Utilizing N-Methyl-2-pyrrolidone (NMP) as the solvent, we combined activated carbon powder samples with polyvinylidene fluoride (PVDF) and carbon black at a weight ratio of 80:10:10 (AC: PVDF: carbon black). The mixtures were stirred with a magnetic stirrer to achieve a homogeneous slurry, and the resulting activated carbon slurry was applied onto a nickel foam current collector. The electrodes were assembled into a symmetric coin cell with a polypropylene surfactant-coated separator (Celgard 3501) and 6 M KOH as electrolyte. Electrochemical assessments were carried out using a potentiostat, specifically the Gamry Reference 3000. These assessments included cyclic voltammetry (CV) and galvanostatic charge–discharge (GCD) to calculate specific capacitance and capacity. Additionally, electrochemical impedance spectroscopy (EIS) was conducted to evaluate impedance characteristics. The specific capacitance obtained from cyclic voltammetry was computed using the following Equation (1) [26]:
C S = 4 I   d v u m Δ V
where u is the scan rate (mV/s), m is the total mass of the two electrodes (g), and ΔV denotes voltage window (V).
We can also calculate specific capacitance from galvanostatic charge–discharge graphs at specific discharge current I, discharge time Δt [2]:
C S = 4 I Δ t m Δ V
From specific capacitance, energy (Wh/kg) and power densities (W/kg) were calculated using the following Equations (3) and (4) given a voltage window V, and discharge time t:
E = 1 2 C s V 2
P = E t

3. Results and Discussion

3.1. Physical Characteristics

The details of the carbonization and activation process for the pecan shell are demonstrated in Figure 1. The emitted gas from carbonization of the pecan shell at a temperature of ~1050 °C may serve as activating agents. These gaseous agents are identified as self-emitted gases for the activation of the carbonization process. The nitrogen (N2) adsorption–desorption isotherms at 77 K of the PSAC of 5, 10, and 15 h of activation time are shown in (Figure 2a). The isotherm curves (Figure 2a) exhibited concave isotherm type I (according to IUPAC), indicating that the PSAC are microporous solids [27]. As P approached P0, the specific surface area of the PSAC increased minimally. The amount of gas adsorbed by PSAC-15 and PSAC-5 is very low, indicating low specific surface area. The BET specific surface areas of PSAC-5, PSAC-10, and PSAC-15 measured were 486.4 m2/g, 1553.6 m2/g, and 538.7 m2/g, respectively (Figure 2c). When activation time was increased from 5 to 10 h, the specific surface area tripled, which is attributed to pore expansion. The resulting decrease in surface area when the activation time changes from 10 to 15 h can be ascribed to the overexpansion of pores leading to breaking of cell walls, reducing pore volume and surface area [22]. The measurement of pore size distribution and specific pore volume (SPV) is calculated by density functional theory (DFT) model as illustrated by Figure 2b. Despite the different activation times, the pore size distribution for all three samples were similar in their high concentration of micropores (width < 2 nm) as seen in Table 1. The PSAC-10 demonstrated the widest pore size distribution as compared to others (PSAC-5 and PSAC-15). Raman spectra (Figure 2d) of the PSACs showed prominent D (1335 cm−1) and G (1590 cm−1) bands. The D band is a result of the disordered or defected carbon structure in the activated pecan shell carbon, while G band represents graphite-like structure.
The ID/IG for PSAC-5, PSAC-10, and PSAC-15 are 1.04, 1.181, and 1.471, respectively. The high degree of disorderness of PSAC-15 can be attributed to its long activation time and is further supported by the emergence of 2D peak (~2600 cm−1), representing the degree of crystallinity and graphene structure [28,29].
The SEM images of the PSAC-10 are presented in Figure 3a,b for investigating the morphology and pore structure of the ACs. A TEM image of the PSAC-10 is shown in Figure 3c,d, where the disordered amorphous carbon structure of PSACs is clearly seen. This result was further confirmed by the diffraction patterns in the inset of Figure 3c and is consistent with the Raman analysis [30]. A uniform distribution of micropores with a pore size of < 2 nm for the PSAC is shown in Figure 3d, which support the pore size distribution result as determined by BET method (Figure 2b).

3.2. Electrochemical Performance

Thorough analysis of the electrochemical characteristics of PSAC involved the application of cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and electrochemical impedance spectroscopy (EIS). These assessments were carried out within a two-electrode coin cell configuration, employing a 6 M KOH electrolyte solution. Only the electrochemical performance of PSAC-10 was evaluated as it exhibited superior physical properties compared to PSAC-5 and PSAC-15. Figure 4a shows cyclic voltammetry at 10 mV/s scan rate in different voltage windows. Figure 4b illustrates the cyclic voltammetry (CV) curves of PSAC-10, showcasing scan rates ranging from 10 to 100 mV/s. The well-defined rectangular characteristics with no redox peaks show the EDLC supercapacitor behavior of PSAC-10. At a high scan rate of 100 mV/s, the electrode material exhibited a slight departure from the ideal rectangular shape with no faradic reaction peaks which indicates the electrode’s ability to facilitate rapid ion transfer and diffusion [31]. The excellent reversibility of the electrode is evident in the diagonal symmetry observed in the rectangular cyclic voltammetry (CV) curve [32,33]. The satisfactory supercapacitor performance of PSAC-10 can be attributed to the balanced ratio of micropores and optimally sized mesopores (2–5 nm) [34].
Galvanostatic charge–discharge profiles (Figure 4c) show typical symmetric linear triangular curves with minimal IR drops at all current densities of 2, 4, 8, and 12 A/g. The minimal IR drop for all current densities reveals PSAC-based electrodes to have low internal resistance. The calculated specific capacitances (Figure 4d) from the discharge curves are 269, 240, 214, 196 F/g at current densities 2, 4, 8, 12 A/g, respectively. The specific capacitance of PSAC-10 was only reduced by 27% when current density is increased from 2 A/g to 12 A/g.
Figure 5a,b show the long-term cyclability of the PSAC-10 for 15,000 cycles at a current density of 2 A/g demonstrating excellent capacity retention of 90% with an average coulombic efficiency of 97%. The high performance of PSAC-10 can be attributed to its high surface area of 1554 m2/g and pore volume of 0.74360 cm3/g (0.44459 cm3/g for micropore and 0.28698 cm3/g mesopores). Consequently, the high percentage of micropore and mesopores makes the material optimal for the efficient adsorption of electrolytic ions i.e., K+ and OH ions [35,36]. It is to be noted that K+ has a bare ionic size of 1.33 Å (0.133 nm) and a hydrated ion size of 3.31 Å (0.331 nm) while OH has a bare ionic size of 1.76 Å (0.176 nm) and a hydrated ion size of 3.00 Å (0.300 nm) [35].
Electrochemical impedance spectroscopy was conducted to measure the internal and overall cell resistance (Figure 5c). The equivalent circuit represents the electrochemical cell (Figure 5c inset). The intercept of x-axis in Nyquist plot represents ohmic resistance (Rs) in the high-frequency range, which arises from electrolyte, current collector, and activated carbon film [37]. The ohmic resistance of ~2 Ω suggests excellent electrical conductivity and good quality of as-fabricated activated carbons. The semi-circle appearing in medium frequency range exemplify RcCc unit attributing to charge transfer resistance (~2 Ω) at the electrode/electrolyte resistance [38]. The straight line in low frequency range is ascribed to capacitance (C).
In comparison with previous studies [9,27,39,40,41,42,43,44,45,46] on biomass-derived activated carbons for supercapacitors, PSAC-10 demonstrated consistently higher energy density and power density calculated at a range of current densities such as 2, 4, 8, and 12 A/g, as shown in Figure 5d. This is due to the unique self-activation process and pecan shell precursor structure. Notably PSAC-10 achieved an energy density of 36 Wh/kg and power density of 2078 W/kg while also maintaining almost all initial capacitance through 15,000 cycles.
Figure 5. (a) Specific capacitance of PSAC-10 supercapacitor over 15,000 charge–discharge cycles at a current density of 2 A/g; (b) galvanostatic charge–discharge profile of PSAC-10 supercapacitor with current density of 2 A/g; (c) electrochemical impedance spectroscopy of Nyquist plot of PSAC-10 supercapacitor with equivalent circuit as inset; (d) Ragone plot of energy and power density values of PSAC-10 in compared with other reported works at 1 V working voltage [9,27,39,40,41,42,43,44,45,46].
Figure 5. (a) Specific capacitance of PSAC-10 supercapacitor over 15,000 charge–discharge cycles at a current density of 2 A/g; (b) galvanostatic charge–discharge profile of PSAC-10 supercapacitor with current density of 2 A/g; (c) electrochemical impedance spectroscopy of Nyquist plot of PSAC-10 supercapacitor with equivalent circuit as inset; (d) Ragone plot of energy and power density values of PSAC-10 in compared with other reported works at 1 V working voltage [9,27,39,40,41,42,43,44,45,46].
Materials 17 03091 g005

4. Conclusions

The as-fabricated pecan shell-derived activated carbon (PSAC) was synthesized by an environmentally benign self-activation process, and the electrochemical performance of PSAC as supercapacitor electrodes was evaluated. The PSAC, fine-tuned under varied activation times of 5, 10, and 15 h, demonstrated outstanding physical characteristics, boasting a high specific surface area (SSA) of 1554 m2/g and a specific pore volume (SPV) of 0.74360 cm3/g. After 10 h of activation, the PSAC-10 exhibits an optimal combination of micro and meso pores, enabling efficient adsorption by offering electroactive sites. This leads to a specific capacitance of 269 F/g at a current density of 2 A/g, accompanied by an impressive cycling stability of around 93% after 15,000 cycles. Furthermore, PSAC-10 maintained 73% of its capacitance at current density 2 A/g when current density was increased to 12 A/g. This straightforward method for creating activated carbon from pecan shells via a novel self-activation method with strong electrochemical capabilities could provide a practical route to environmentally friendly and easily scalable supercapacitor production.

Author Contributions

Conceptualization, S.J.Z., S.Q.S.; methodology, M.D.P., L.M.S., E.C., S.Q.S. and W.C.; software, M.D.P., L.M.S., E.C.; validation, M.D.P., L.M.S., E.C., S.Q.S. and W.C.; formal analysis, S.J.Z., M.D.P., L.M.S., E.C.; investigation, S.J.Z., M.D.P., L.M.S., E.C., S.Q.S. and W.C.; resources, S.Q.S. and W.C.; data curation, S.J.Z. and M.D.P.; writing—original draft preparation, S.J.Z.; writing—review and editing, S.J.Z., M.D.P., L.M.S., S.Q.S. and W.C.; visualization, S.J.Z., M.D.P., L.M.S., S.Q.S. and W.C.; supervision, S.Q.S. and W.C.; project administration, S.Q.S. and W.C.; funding acquisition, S.Q.S. and W.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Winter, M.; Brodd, R.J. What Are Batteries, Fuel Cells, and Supercapacitors? Chem. Rev. 2005, 105, 1021. [Google Scholar] [CrossRef]
  2. Simon, P.; Gogotsi, Y. Materials for Electrochemical Capacitors. Nat. Mater. 2008, 7, 845–854. [Google Scholar] [CrossRef] [PubMed]
  3. Miller, J.R.; Simon, P. Electrochemical Capacitors for Energy Management. Sci. Mag. 2008, 321, 651–652. [Google Scholar] [CrossRef] [PubMed]
  4. Devillers, N.; Jemei, S.; Péra, M.C.; Bienaimé, D.; Gustin, F. Review of Characterization Methods for Supercapacitor Modelling. J. Power Sources 2014, 246, 596–608. [Google Scholar] [CrossRef]
  5. Kim, B.K.; Sy, S.; Yu, A.; Zhang, J. Electrochemical Supercapacitors for Energy Storage and Conversion. In Handbook of Clean Energy Systems; Wiley: Hoboken, NJ, USA, 2015. [Google Scholar]
  6. Cao, Y.; Wang, K.; Wang, X.; Gu, Z.; Fan, Q.; Gibbons, W.; Shrestha, M. Hierarchical Porous Activated Carbon for Supercapacitor Derived from Corn Stalk Core by Potassium Hydroxide Activation. Electrochim. Acta 2016, 212, 839–847. [Google Scholar] [CrossRef]
  7. Jisha, M.R.; Hwang, Y.J.; Shin, J.S.; Nahm, K.S.; Kumar, T.P.; Karthikeyan, K.; Stephan, A.M. Electrochemical Characterization of Supercapacitors Based on Carbons Derived from Coffee Shells. Mater. Chem. Phys. 2009, 115, 33–39. [Google Scholar] [CrossRef]
  8. Sun, W.; Lipka, S.M.; Swartz, C.; Williams, D.; Yang, F. Hemp-Derived Activated Carbons for Supercapacitors. Carbon 2016, 103, 181–192. [Google Scholar] [CrossRef]
  9. Qian, W.; Sun, F.; Xu, Y.; Qiu, L.; Liu, C.; Wang, S.; Yan, F. Human Hair-Derived Carbon Flakes for Electrochemical Supercapacitors. Energy Environ. Sci. 2014, 7, 379–386. [Google Scholar] [CrossRef]
  10. Yin, J.; Zhu, Y.; Yue, X.; Wang, L.; Zhu, H.; Wang, C. From Environmental Pollutant to Activated Carbons for High-Performance Supercapacitors. Electrochim. Acta 2016, 201, 96–105. [Google Scholar] [CrossRef]
  11. Preston, C.M.; Sayer, B.G. What’s in a Nutshell: An Investigation of Structure by Carbon-13 Cross-Polarization Magic-Angle Spinning Nuclear Magnetic Resonance Spectroscopy. J. Agric. Food Chem. 1992, 40, 206–210. [Google Scholar] [CrossRef]
  12. Xu, J.; Gao, Q.; Zhang, Y.; Tan, Y.; Tian, W.; Zhu, L.; Jiang, L. Preparing Two-Dimensional Microporous Carbon from Pistachio Nutshell with High Areal Capacitance as Supercapacitor Materials. Sci. Rep. 2014, 4, 5545. [Google Scholar] [CrossRef] [PubMed]
  13. Bansode, R.R.; Losso, J.N.; Marshall, W.E.; Rao, R.M.; Portier, R.J. Adsorption of Metal Ions by Pecan Shell-Based Granular Activated Carbons. Bioresour. Technol. 2003, 89, 115–119. [Google Scholar] [CrossRef] [PubMed]
  14. Bansode, R.R.; Losso, J.N.; Marshall, W.E.; Rao, R.M.; Portier, R.J. Adsorption of Volatile Organic Compounds by Pecan Shell-and Almond Shell-Based Granular Activated Carbons. Bioresour. Technol. 2003, 90, 175–184. [Google Scholar] [CrossRef] [PubMed]
  15. Toles, C.A.; Marshall, W.E.; Johns, M.M. Granular Activated Carbons from Nutshells for the Uptake of Metals and Organic Compounds. Carbon 1997, 35, 1407–1414. [Google Scholar] [CrossRef]
  16. Yang, K.; Peng, J.; Srinivasakannan, C.; Zhang, L.; Xia, H.; Duan, X. Preparation of High Surface Area Activated Carbon from Coconut Shells Using Microwave Heating. Bioresour. Technol. 2010, 101, 6163–6169. [Google Scholar] [CrossRef] [PubMed]
  17. Xu, B.; Chen, Y.; Wei, G.; Cao, G.; Zhang, H.; Yang, Y. Activated Carbon with High Capacitance Prepared by NaOH Activation for Supercapacitors. Mater. Chem. Phys. 2010, 124, 504–509. [Google Scholar] [CrossRef]
  18. Elmouwahidi, A.; Zapata-Benabithe, Z.; Carrasco-Marín, F.; Moreno-Castilla, C. Activated Carbons from KOH-Activation of Argan (Argania spinosa) Seed Shells as Supercapacitor Electrodes. Bioresour. Technol. 2012, 111, 185–190. [Google Scholar] [CrossRef] [PubMed]
  19. Boyjoo, Y.; Cheng, Y.; Zhong, H.; Tian, H.; Pan, J.; Pareek, V.K.; Liu, J. From Waste Coca Cola® to Activated Carbons with Impressive Capabilities for CO2 Adsorption and Supercapacitors. Carbon 2017, 116, 490–499. [Google Scholar] [CrossRef]
  20. Liu, Q.S.; Zheng, T.; Wang, P.; Guo, L. Preparation and Characterization of Activated Carbon from Bamboo by Microwave-Induced Phosphoric Acid Activation. Ind. Crops Prod. 2010, 31, 233–238. [Google Scholar] [CrossRef]
  21. Karagöz, S.; Tay, T.; Ucar, S.; Erdem, M. Activated Carbons from Waste Biomass by Sulfuric Acid Activation and Their Use on Methylene Blue Adsorption. Bioresour. Technol. 2008, 99, 6214–6222. [Google Scholar] [CrossRef] [PubMed]
  22. Xia, C.; Shi, S.Q. Self-Activation for Activated Carbon from Biomass: Theory and Parameters. Green Chem. 2016, 18, 2063–2071. [Google Scholar] [CrossRef]
  23. Bommier, C.; Xu, R.; Wang, W.; Wang, X.; Wen, D.; Lu, J.; Ji, X. Self-Activation of Cellulose: A New Preparation Methodology for Activated Carbon Electrodes in Electrochemical Capacitors. Nano Energy 2015, 13, 709–717. [Google Scholar] [CrossRef]
  24. Xia, C.; Shi, S.Q. Self-Activation Process to Fabricate Activated Carbon from Kenaf. Wood Fiber Sci. 2016, 48, 62–69. [Google Scholar]
  25. Xia, C.; Kang, C.; Patel, M.D.; Cai, L.; Gwalani, B.; Banerjee, R.; Choi, W. Pine Wood Extracted Activated Carbon through Self-Activation Process for High-Performance Lithium-Ion Battery. ChemistrySelect 2016, 1, 4000–4007. [Google Scholar] [CrossRef]
  26. Stoller, M.D.; Ruoff, R.S. Best Practice Methods for Determining an Electrode Material’s Performance for Ultracapacitors. Energy Environ. Sci. 2010, 3, 1294–1301. [Google Scholar] [CrossRef]
  27. Biswal, M.; Banerjee, A.; Deo, M.; Ogale, S. From Dead Leaves to High Energy Density Supercapacitors. Energy Environ. Sci. 2013, 6, 1249–1259. [Google Scholar] [CrossRef]
  28. Escribano, R.; Sloan, J.J.; Siddique, N.; Sze, N.; Dudev, T. Raman Spectroscopy of Carbon-Containing Particles. Vib. Spectrosc. 2001, 26, 179–186. [Google Scholar] [CrossRef]
  29. Yun, Y.S.; Cho, S.Y.; Shim, J.; Kim, B.H.; Chang, S.J.; Baek, S.J.; Jin, H.J. Microporous Carbon Nanoplates from Regenerated Silk Proteins for Supercapacitors. Adv. Mater. 2013, 25, 1993–1998. [Google Scholar] [CrossRef] [PubMed]
  30. Marsh, H.; Reinoso, F.R. Activated Carbon; Elsevier: Amsterdam, The Netherlands, 2006. [Google Scholar]
  31. Stoller, M.D.; Park, S.; Zhu, Y.; An, J.; Ruoff, R.S. Graphene-based ultracapacitors. Nano Lett. 2008, 8, 3498–3502. [Google Scholar] [CrossRef] [PubMed]
  32. Wei, L.; Yushin, G. Electrical Double Layer Capacitors with Activated Sucrose-Derived Carbon Electrodes. Carbon 2011, 49, 4830–4838. [Google Scholar] [CrossRef]
  33. Xu, B.; Wu, F.; Chen, R.; Cao, G.; Chen, S.; Zhou, Z.; Yang, Y. Highly Mesoporous and High Surface Area Carbon: A High Capacitance Electrode Material for EDLCs with Various Electrolytes. Electrochem. Commun. 2008, 10, 795–797. [Google Scholar] [CrossRef]
  34. Chmiola, J.; Yushin, G.; Gogotsi, Y.; Portet, C.; Simon, P.; Taberna, P.L. Anomalous Increase in Carbon Capacitance at Pore Sizes Less than 1 Nanometer. Science 2006, 313, 1760–1763. [Google Scholar] [CrossRef] [PubMed]
  35. Zhong, C.; Deng, Y.; Hu, W.; Qiao, J.; Zhang, L.; Zhang, J. A Review of Electrolyte Materials and Compositions for Electrochemical Supercapacitors. Chem. Soc. Rev. 2015, 44, 7484–7539. [Google Scholar] [CrossRef] [PubMed]
  36. Frackowiak, E. Carbon Materials for Supercapacitor Application. Phys. Chem. Chem. Phys. 2007, 9, 1774–1785. [Google Scholar] [CrossRef] [PubMed]
  37. Kötz, R.; Hahn, M.; Gallay, R. Temperature Behavior and Impedance Fundamentals of Supercapacitors. J. Power Sources 2006, 154, 550–555. [Google Scholar] [CrossRef]
  38. Wang, H.; Hao, Q.; Yang, X.; Lu, L.; Wang, X. Graphene Oxide Doped Polyaniline for Supercapacitors. Electrochem. Commun. 2009, 11, 1158–1161. [Google Scholar] [CrossRef]
  39. Zhang, Y.; Jia, M.; Gao, H.; Yu, J.; Wang, L.; Zou, Y.; Zhao, Y. Porous Hollow Carbon Spheres: Facile Fabrication and Excellent Supercapacitive Properties. Electrochim. Acta 2015, 184, 32–39. [Google Scholar] [CrossRef]
  40. Yu, B.; Chang, Z.; Wang, C. The Key Pre-Pyrolysis in Lignin-Based Activated Carbon Preparation for High Performance Supercapacitors. Mater. Chem. Phys. 2016, 181, 187–193. [Google Scholar] [CrossRef]
  41. Yin, L.; Chen, Y.; Zhao, X.; Hou, B.; Cao, B. 3-Dimensional Hierarchical Porous Activated Carbon Derived from Coconut Fibers with High-Rate Performance for Symmetric Supercapacitors. Mater. Des. 2016, 111, 44–50. [Google Scholar] [CrossRef]
  42. Rufford, T.E.; Hulicova-Jurcakova, D.; Zhu, Z.; Lu, G.Q. Nanoporous Carbon Electrode from Waste Coffee Beans for High Performance Supercapacitors. Electrochem. Commun. 2008, 10, 1594–1597. [Google Scholar] [CrossRef]
  43. Raymundo-Piñero, E.; Cadek, M.; Béguin, F. Tuning Carbon Materials for Supercapacitors by Direct Pyrolysis of Seaweeds. Adv. Funct. Mater. 2009, 19, 1032–1039. [Google Scholar] [CrossRef]
  44. Rufford, T.E.; Hulicova-Jurcakova, D.; Khosla, K.; Zhu, Z.; Lu, G.Q. Microstructure and Electrochemical Double-Layer Capacitance of Carbon Electrodes Prepared by Zinc Chloride Activation of Sugar Cane Bagasse. J. Power Sources 2010, 195, 912–918. [Google Scholar] [CrossRef]
  45. Han, Y.; Shen, N.; Zhang, S.; Li, D.; Li, X. Fish Gill-Derived Activated Carbon for Supercapacitor Application. J. Alloys Compd. 2017, 694, 636–642. [Google Scholar] [CrossRef]
  46. Feng, H.; Hu, H.; Dong, H.; Xiao, Y.; Cai, Y.; Lei, B.; Zheng, M. Hierarchical Structured Carbon Derived from Bagasse Wastes: A Simple and Efficient Synthesis Route and Its Improved Electrochemical Properties for High-Performance Supercapacitors. J. Power Sources 2016, 302, 164–173. [Google Scholar] [CrossRef]
Figure 1. Schematic and the corresponding temperature vs. time graph outlining the self-activation process of the pecan shell. The middle segment represents dwelling temperature and time.
Figure 1. Schematic and the corresponding temperature vs. time graph outlining the self-activation process of the pecan shell. The middle segment represents dwelling temperature and time.
Materials 17 03091 g001
Figure 2. (a) Nitrogen adsorption–desorption for PSAC-5, 10, and 15; (b) pore size distribution for PSAC-5, PSAC-10, PSAC-15; (c) surface area comparisons between PSAC-5, PSAC-10, and PSAC-15; (d) Raman spectra of PSAC-5, PSAC-10, PSAC-15.
Figure 2. (a) Nitrogen adsorption–desorption for PSAC-5, 10, and 15; (b) pore size distribution for PSAC-5, PSAC-10, PSAC-15; (c) surface area comparisons between PSAC-5, PSAC-10, and PSAC-15; (d) Raman spectra of PSAC-5, PSAC-10, PSAC-15.
Materials 17 03091 g002
Figure 3. SEM and TEM Images of PSAC-10: (a,b) SEM image of PSAC-10 interconnected pore structure; (c) 100 nm width TEM displays amorphous structure that is confirmed with inset; (d) 10 nm width TEM showing micropore structure.
Figure 3. SEM and TEM Images of PSAC-10: (a,b) SEM image of PSAC-10 interconnected pore structure; (c) 100 nm width TEM displays amorphous structure that is confirmed with inset; (d) 10 nm width TEM showing micropore structure.
Materials 17 03091 g003
Figure 4. (a) Cyclic voltammetry of PSAC-10 across voltage windows of −0.5–0.5 and 0–1.0 V; (b) cyclic voltammetry of PSAC-10 supercapacitor in 6 M KOH with voltage window of 0–1.0 V and scan rates of 10 mV/s, 20 mV/s, 50 mV/s, and 100 mV/s; (c) galvanostatic charge–discharge profile of PS-10 supercapacitor with current density of 2 A/g, 4 A/g, 8 A/g, and 12 A/g in a voltage range of 0–1.0 V; (d) specific capacitance plotted against current density calculated from GCD profiles.
Figure 4. (a) Cyclic voltammetry of PSAC-10 across voltage windows of −0.5–0.5 and 0–1.0 V; (b) cyclic voltammetry of PSAC-10 supercapacitor in 6 M KOH with voltage window of 0–1.0 V and scan rates of 10 mV/s, 20 mV/s, 50 mV/s, and 100 mV/s; (c) galvanostatic charge–discharge profile of PS-10 supercapacitor with current density of 2 A/g, 4 A/g, 8 A/g, and 12 A/g in a voltage range of 0–1.0 V; (d) specific capacitance plotted against current density calculated from GCD profiles.
Materials 17 03091 g004
Table 1. Table of BET specific surface area (SSA), pore volume, and pore size distribution of the PSACs.
Table 1. Table of BET specific surface area (SSA), pore volume, and pore size distribution of the PSACs.
Sample BET SSA SPV (cm3/g)
(m2g−1)MicroporeMesoporesMacroporeTotal
PSAC-5486.10.20433 (86.8%)0.02378
(10.1%)
0.00739
(3.14%)
0.23550
PSAC-101553.60.44459
(59.8%)
0.28698
(38.6%)
0.01203
(1.62%)
0.74360
PSAC-15540.80.19096
(60.7%)
0.11693
(37.2%)
0.00681
(2.2%)
0.31470
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

Zou, S.J.; Patel, M.D.; Smith, L.M.; Cha, E.; Shi, S.Q.; Choi, W. Pecan Shell-Derived Activated Carbon for High-Electrochemical Performance Supercapacitor Electrode. Materials 2024, 17, 3091. https://doi.org/10.3390/ma17133091

AMA Style

Zou SJ, Patel MD, Smith LM, Cha E, Shi SQ, Choi W. Pecan Shell-Derived Activated Carbon for High-Electrochemical Performance Supercapacitor Electrode. Materials. 2024; 17(13):3091. https://doi.org/10.3390/ma17133091

Chicago/Turabian Style

Zou, Sarah J., Mumukshu D. Patel, Lee M. Smith, Eunho Cha, Sheldon Q. Shi, and Wonbong Choi. 2024. "Pecan Shell-Derived Activated Carbon for High-Electrochemical Performance Supercapacitor Electrode" Materials 17, no. 13: 3091. https://doi.org/10.3390/ma17133091

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