Next Article in Journal
Comprehensive Evaluation and Analysis of Human Settlements’ Suitability in the Yangtze River Delta Based on Multi-Source Data
Next Article in Special Issue
Effects of Magnetic Biochar Addition on Mesophilic Anaerobic Digestion of Sewage Sludge
Previous Article in Journal
Aircraft Noise Reduction Strategies and Analysis of the Effects
Previous Article in Special Issue
Co-Thermal Oxidation of Lignite and Rice Straw for Synthetization of Composite Humic Substances: Parametric Optimization via Response Surface Methodology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation and Electrochemical Performance of Bio-Oil-Derived Hydrochar as a Supercapacitor Electrode Material

1
Joint International Research Laboratory of Biomass Energy and Materials, Co-Innovation Center of Efficient Processing and Utilization of Forest Resources, College of Materials Science and Engineering, Nanjing Forestry University, Nanjing 210037, China
2
School of Energy and Power Engineering, Jiangsu University, Zhenjiang 212013, China
3
State Key Laboratory of High-Efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University, Yinchuan 750021, China
*
Authors to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2023, 20(2), 1355; https://doi.org/10.3390/ijerph20021355
Submission received: 15 December 2022 / Revised: 10 January 2023 / Accepted: 10 January 2023 / Published: 11 January 2023
(This article belongs to the Special Issue New Advances in Biomass Conversion and the Environmental Implications)

Abstract

:
The rapid consumption of fossil energy and the urgent demand for sustainable development have significantly promoted worldwide efforts to explore new technology for energy conversion and storage. Carbon-based supercapacitors have received increasing attention. The use of biomass and waste as a carbon precursor is environmentally friendly and economical. In this study, hydrothermal pretreatment was used to synthetize coke from bio-oil, which can create a honeycomb-like structure that is advantageous for electrolyte transport. Furthermore, hydrothermal pretreatment, which is low in temperature, can create a low graphitization degree which can make heteroatom introduction and activation easier. Then, urea and KOH were used for doping and activation, which can improve conductivity and capacitance. Compared with no heteroatom and activation hydrothermal char (HC) (58.3 F/g at 1 A/g), the prepared carbon material nitrogen doping activated hydrothermal carbon (NAHC1) had a good electrochemical performance of 225.4 F/g at 1 A/g. The specific capacitance of the prepared NAHC1 was improved by 3.8 times compared with that of HC.

1. Introduction

The rapid consumption of fossil energy and the urgent demand for sustainable development have greatly promoted worldwide efforts to explore new technology for energy conversion and storage [1]. Supercapacitors, which are energy storage devices, have many advantages compared with traditional storage devices, such as high power density, long lifecycle, fast charge and discharge performance, etc. Therefore, supercapacitors have attracted increasing interest among many researchers. The storage types of supercapacitors include (1) electric double-layer capacitance; (2) and pseudo capacitance [2]. The mechanism of electric double-layer capacitance is that the ion is stored based on physical electrostatic adsorption without any chemical reaction during charging and discharging [3]. The mechanism of pseudo capacitance is that charge storage occurs via a fast reversible redox reaction on or near the surface [4]. For electric double-layer capacitors (EDLCs), the carbon materials are widely applied [5]. Biomass and waste [6,7,8], including onions, houttuynia, lotus stalks, bio-oil, etc., are significant sources or precursors to carbon materials due to their high carbon content and low cost. For example, some researchers [9] use yeast as a carbon precursor to prepare electrode material, which has good electrochemistry performance that has 316 F g−1 at 1 A g−1.
However, the performance of carbon derived from biomass is significantly low for its complex structure and chemical composition. Many researchers seek methods to enhance the electrochemical performance of the materials. Widely used means of modification involve activation and doping.
The activation treatment of materials is the most effective way to enhance performance by raising materials’ specific surface area. Methods of activation treatment are usually divided into two kinds, namely chemical activation and physical activation [10]. Chemical activation is the most convenient and effective method, which generally uses potassium hydroxide (KOH) [11], sodium hydroxide (NaOH), salt, etc. Chemical activation can produce a high specific surface area and porous structure. Wang et al. [12] employed KOH as an activator to activate carbon from Metaplexis japonica. The sample has an ultrahigh surface area of about 2210 m2 g−1 and high specific capacitance of 287 F g−1 at 1.0 A g−1. Wang et al. [13] used a mixture of NaCl and ZnCl2 activators to create a surface area of about 484 m2 g−1 and the sample had a specific capacitance of 293 F g−1. Due to the great electrode performance of the heteroatom doping carbon, some researchers have applied doping to enhance electrochemical performance. Nitrogen (N) [14], sulfur (S) [15], and boron (B) [16] are generally used as the doping element. There are two methods of doping heteroatoms: one is self-doping and the other is chemical doping. Self-doping [17] involves the heteroatoms of electrode materials being introduced by raw materials. Chemical doping [18] involves the heteroatoms of electrode materials being introduced by using other sources. Owing to the size of nitrogen compared with that of carbon, nitrogen doping is the most common method in the heteroatoms doping of electrode materials [19]. Because nitrogen heteroatoms in porous carbon can improve surface wettability and electronic conductivity and induce pseudo capacitance, they thus promote the capacitance of supercapacitors [20]. Many researchers have employed nitrogen doping to improve the electrochemical performance of carbon materials. For example, Wu et al. [21] used lilac and lotus seedpods as carbon precursors and N2 nonthermal plasma generated by the dielectric barrier discharge (DBD) to finally synthesize nitrogen doping carbon materials which had a great electrochemical performance of 342.5 F/g at 0.5 A/g.
Bio-oil is the important product of biomass pyrolysis, and contains a large number of aromatic hydrocarbons and polycyclic carbon compounds [22], which can be easily polymerized and aromatized to generate carbons with heat treatment. Thus, bio-oil is a desirable carbon precursor to synthesize electrode materials. Compared with biomass, bio-oil shows obvious advantages such as low ash content [23] and liquid phase. Accordingly, electrode material derived from bio-oil have better stability and easy modification. Additionally, bio-oil shows high oxygen content. Oxygen content is advantageous to modify the surface electron properties of carbon materials and improve the electrode wettability to increase ion-accessible surface areas [24], which can enhance solid–liquid interface reaction and thus improve the electrochemical performance. Therefore, it can be concluded that bio-oil-derived carbon is a promising choice for supercapacitor electrode material. Prior to actual application, the production of bio-oil-derived carbon is the key. Some researchers [25,26] have successfully used hydrothermal carbonization (HTC) to upgrade bio-oil into a solid fuel, and it was found that HTC was conducive to enhancing carbon content in hydrochar and removing the unstable component in bio-oil through dehydration and decarboxylation reaction [25]. Furthermore, compared with the conventional carbonization method (i.e., pyrolysis or high-temperature carbonization), HTC shows incomparable advantages such as low energy consumption due to its mild reaction conditions, high solid product yield [26], and high oxygen content functional groups on solid surface. Therefore, bio-oil-derived hydrochar might be an attractive supercapacitor electrode material. However, the related study on preparation and electrochemical performance of bio-oil-derived hydrochar as a supercapacitor electrode material was almost a gap.
In this work, a nitrogen doping, porous bio-oil-derived carbon electrode materials were efficiently synthesized. Firstly, bio-oil was treated by hydrothermal pretreatment. Subsequently, nitrogen was introduced into hydrochar using urea. The nitrogen doping carbons were activated by KOH. The morphology of the prepared samples was charactered by SEM. The surface area and pore structure of the samples were analyzed by BET. The graphitization degree of the samples was characterized by XRD and Raman. The doping nitrogen type of the sample was characterized by XPS. The electrochemical performance was characterized by electrochemical workstation at 6 M KOH electrolyte. The final synthesized sample had a good electrochemical performance of 225.4 F/g at 1 A/g.

2. Materials and Methods

2.1. Hydrothermal Pretreatment of Bio-Oil

The feedstock of this study was fast pyrolysis bio-oil (provided by Biomass Technology Group, Enschede, The Netherlands. An amount of 60 g of fast pyrolysis bio-oil (FPBO) was directly poured into a stainless steel kettle. The reactant was subsequently heated to 250 °C at 300 r/min under argon (Ar) atmosphere and then kept at 250 °C for 8 h. The product was cooled to room temperature and then washed by ethanol 5–6 times, which was denoted as HC. The mass yield of HC was about 28.3%.

2.2. Synthesis of N-Doped Carbon from HC

The N-doped carbon was synthesized by carbonizing a mixture of urea and HC. In detail, 2 g HC and 2 g urea were mixed in an agate mortar. The mixture was placed on a quartz boat. Afterwards, the quartz boat was placed into a tubular furnace. The reaction temperature was 700 °C for 1 h with a heating rate of 5 °C/min. Argon (Ar) was served as the carrier gas in the tubular furnace. The product was cooled to room temperature. The product was washed by deionized water 5–6 times. The product was subsequently was dried at 105 °C for 24 h. The product was named NHC1. Similarly, the as-prepared N-doped carbon was named NHC2, NHC3.

2.3. Synthesis of N-Doped Porous Carbon from NHC

N-doped porous carbon NHCs were activated by KOH. In detail, 3 g of KOH was dissolved in deionized water and then 3 g of NHC was added into the solution. The mixed solution was stirred for 2 h at room temperature. After being stirred, the mixed solution was dried at 105 °C for 24 h. the mixed solid powder was placed into the tubular furnace. The reaction temperature was 700 °C for 1 h with a heating rate of 5 °C/min. Argon (Ar) was served as the carrier gas in the tubular furnace. The product was cooled to room temperature. The product was first washed with 0.5 M HCl and then washed with deionized water for 5–6 times. The product was subsequently dried at 105 °C for 24 h. The product was named NAHC1, NAHC2, NAHC3.

2.4. Characterization

Nitrogen adsorption–desorption isotherm measurements were used to analyze the pore structure of samples (BSD-PM4). The surface area and pore size distribution were determined by using the Brunauer–Emmett–Teller theory and the nonlocal density functional theory method, respectively. The morphologies, microstructures, and crystallinity of the samples were analyzed by Quanta 200 scanning electron microscopy (SEM, FEI, Hillsboro, Ultima, IV, USA), X-ray diffraction (XRD, Rigaku, Tokyo, Japan), AXIS UltraDLD X-ray photoelectron spectroscopy (XPS, Shimadzu, Kyoto, Japan), and DXR532 Raman scattering spectroscopy (Raman, ThemorFisher, Waltham, MA, USA )

2.5. Supercapacitor Electrode Preparation and Assembly

The prepared samples were mixed with a conductive carbon black and PTFE emulsion (mass ratio = 8:1:1) in alcohol. The prepared slurry was applied evenly on circular nickel foam to form an electrode after drying [27].

2.6. Electrochemical Measurements

The electrochemical performance of the samples electrode was tested by a three-electrode system. In a typical three-electrode system, the as-prepared electrode, Pt foil and Hg/HgO were used as the working electrode, counter electrode, and reference electrode, respectively. A total of 6 M KOH was used as electrolyte. An electrochemical working station was used for the cyclic voltammetry (CV), galvanostatic charge–discharge (GCD), and impedance measurements.
The specific capacitance of a single electrode was calculated according to the galvanostatic charge–discharge curve using Equation (1) for the three-electrode system [27]:
C m = I Δ t m Δ V
where Cm is the specific capacitance (F/g), I (A) is the discharge current, Δ t (s) is the discharge time, m (g) is the total mass of active materials in both electrodes, and ΔV (V) is the potential change within the discharge time.

3. Results and Discussions

3.1. Morphology and Pore Structure

The SEM images of the samples were shown in Figure 1. All images of the samples presented a sheet structure. Due to high pressure and polarity of water in bio-oil, the bio-oil precursor was transformed to honeycomb-like coke during the hydrothermal process. Honeycomb-like carbon materials are good for the transmission of electrolytes [28]. It was observed that the sheet structure of NAHC1 was destroyed as a result of KOH activation. The activation effect of KOH can improve a specific surface area of the sample, while also making the sheet collapse and furthermore, causing a block in electrolyte transmission. It can be seen in Figure 1a–c that the surface of NHCs were slightly destroyed. It can be inferred that the high temperature and urea pressure vapor destroyed the surface of the sample during the urea vaporization process. After KOH activation, the destruction obviously increased, as seen in Figure 1d,f. With the ratio of urea increasing, the destruction was more obvious, which indicates that the doping of nitrogen would decrease the graphitization degree; thus, the effect of KOH activation was better.
The N2 adsorption–desorption measurement was conducted to analyze the porous structure. The nitrogen adsorption–desorption isotherms of NHC1 and NAHC1 were shown in Figure 2. The nitrogen adsorption–desorption isotherms of NHCs show a type-IV [29] isothermal with low Sbet of 15–16 m2/g, which would seriously affect electrochemical performance. Since the samples were activated by KOH, the pore structure of samples was obviously improved. The nitrogen adsorption–desorption isotherms of all NAHCs were shown a combination of type I, II, and IV [30], with an obvious hysteresis loop at a relative pressure range of 0.4 and 1, which indicates the existence of a hierarchical pore structure. It was confirmed that micropore and mesopore, especially mesopores, can improve electrochemical performance [31]. The detailed pore structure parameter was shown in Table 1. NAHC1 had the largest Smeso (229.2837 m2/g). Although NAHC1 did not have the largest SBet or nitrogen content, the electrochemical performance of NAHC1 was the best, as discussed in the following section.

3.2. Carbon Structure and Elemental Composition

Figure 3a shows the XRD pattern of the NHCs and NAHCs. it shows two wide peaks located at ~24° and ~43° in Figure 3a, which assigns the graphitic crystallite plane (002) and crystallite plane (100) [32]. Compared with NHCs, the peak (002) of NAHCs was broader, which revealed that the graphitization degree of NAHCs became weaker than that of NHCs. Thus, it can be inferred that the activation treatment reduces the graphitization degree of NHCs, which can be proven by Raman spectroscopy.
The Raman spectroscopy was applied to analyze the graphitization of the samples. Two signals, 1350 cm−1 and 1580 cm−1, can be observed in Figure 3b, which represents the D band and G band [33], respectively. These characteristic D and G bands of the carbon materials represent disordered structures and sp2-bonded carbon atoms [34] (E2g mode of 2D graphite), respectively. The intensity ratio of the D/G bands (ID/IG) was used to characterize the degree of defects [35]. The higher ID/IG indicated more defects in the samples. More defects mean a better electrochemical performance. The exhibited ratio ID/IG of NHC1, NAHC1, NHC2, NAHC2, NHC3, and NAHC3 was 1.286, 1.472, 1.320, 1.510, 1.279, and 1.568. The NHCs samples exhibited a lower ratio of ID/IG than the NAHCs, which implied that the effect of the KOH activation treatment would reduce the degree of the graphitization. With the increasing ratio of nitrogen, the ratio of ID/IG of NAHCs also increased. The highest degree of defects was detected in the NAHC3 sample. The results suggest that: (1) a KOH activation treatment and nitrogen doping treatment would increase the ratio of ID/IG of the material; (2) and with the increasing content of nitrogen, the ratio of ID/IG also increased.
X-ray photoelectron spectroscopy was applied to investigate the modality of nitrogen doping. Table 2 shows all the samples’ elemental compositions. Due to the effect of KOH activation, the nitrogen contents of all samples were decreased. It was inferred that KOH activation had essential damage to the nitrogen-containing structure. Figure 4a,c shows the high-resolution XPS C 1s of the NHC1 and NAHC1. The C 1s spectrum can be divided into four fitted peaks, 284.5, 285.7, 287.7, and 289.6 eV [36], which represent C-C/C=C, C-N, C=O, and O-C=O, respectively. The N 1s spectrum form in Figure 4b,d can be divided into four and three fitted peaks corresponding to NHC1 and NAHC1, respectively. The N 1s spectrum of NHC1 contains four peaks, 398.1, 400.1, 401.8, and 403.5 eV [34], which are assigned to the pyridinic-N(N-6), pyrrolic-N(N-5), quaternary-N(N-Q), and pyridinic-N-oxide(N-X), respectively. However, the N 1s spectrum of NAHC1 only contains three peaks, which are N-6, N-5, and N-Q. Compared with the N 1s of NHC, it can be seen that the intensity of NAHC’s N-X was disappeared and the intensity of N-5, N-6, and N-Q was enhanced. Due to the existence of π-conjugated skeletons, N-5, N-6, and N-Q were stable in the activation treatment of NHC [37]. Moreover, N-X would be destroyed in the activation treatment. the previous research proved that N-5 and N-6 were expected to provide active sites to enhance electrochemically performance [38]. Furthermore, the previous work proved that N-Q can enhance the wettability of the electrode and N-5 and N-6 can improve electron mobility [39].

3.3. Electrochemical Performance

To characterize the electrochemical performance of NHCs and NAHCs, CV and GCD curves were measured in an aqueous electrolyte solution of 6 M KOH using a three-electrode system [30]. CV curves were tested 5 cycles before it was collected, and each sample was tested 3 times. As shown in Figure 5a, the CV curves at 10 mV/s of NAHCs show quasi-rectangular shapes, which revealed a quick electrochemical response and dominant electric double-layer capacitance behavior [40]. A weak and broad peak can be observed, which may be derived from the reversible redox reaction due to the existence of nitrogen [41]. The integral area of the CV curve indicates the electrode’s ion adsorption capacity. The integral area of NHCs’ CV curve shows a low size, which indicates a weak electrode’s ion adsorption. It can be inferred that a low specific surface area would be adverse to an electrode’s ion adsorption. Nevertheless, the integral area of the NAHCs’ CV curve was large and it was increased with the decrease in doping ratio. By the above BET measurement, it was found that the high specific surface area can improve an electrode’s ion adsorption. The CV curves at 100 mV/s show a similar trend in Figure 5b; although the peak was more obvious, the curves kept a quasi-rectangular shape at 100 mV/s.
Capacitive performance was evaluated by a GCD measurement. The GCD curves of all samples displayed a nearly straight line and are symmetrical in Figure 5c, which can imply that the samples represent a dominant electric double-layer capacitance (EDLCs) behavior and a good charge–discharge process. The time of the charge–discharge process reflected the capacitance of samples by the above formula, the specific capacitance can be calculated. Among NHCs, NHC2 had the highest specific capacitance, which was 82.8 F/g at 1 A/g. NHC1 and NHC3 had close specific capacitances, which were 65.2 F/g and 63.2 F/g at 1 A/g, respectively. Through KOH activation, specific capacitances had a huge enhancement. NAHC1 had the highest specific capacitance, which was 210.5 F/g at 1 A/g. The specific capacitance of NAHC1 was 3.2 times higher than NHC1, which illustrated that enhancement of a specific surface area would improve the electrochemical performance. The GCD curves of all samples at 10 A/g were shown in Figure 5d. NAHC1 maintained good electrochemical performance, which was 151 F/g. The specific capacitance of other samples, especially NHCs, seriously decreased.
An EIS measurement was employed to study the resistance and capacitance of samples. At the low frequency region in Figure 5e, it can be seen that the more inclined a line is close to the theoretical vertical line, the better the capacitance performance of the samples are. In Figure 5e, the lines of NAHCs are closer to the theoretical vertical line [40], especially NAHC3; therefore, NAHCs have better capacitance. The lines of NHCs were not good even in not straight lines, which may be due to the poor pore structure of NHCs. At the high frequency region shown in Figure 5f, NAHCs and NHCs all have a low internal resistance (Rs) (0.648–0.704 Ohm) value and charge transfer resistance (Rct) due to the doping of the nitrogen which can improve hydrophilicity and the transmission performance of electrolytes in the materials.
To further explore the performance of NAHC1, CV and GCD were employed in Figure 6. In Figure 6a, all CV curves from 10 mV/s to 100 mV/s were quasi-rectangular in shapes, which indicated that NAHC1had EDLC behavior from 10 mV/s to 100 mV/s. With the sweep rate increasing, the peak of NAHC1 was more obvious due to the enhancement of pseudo capacitance. In Figure 6b, the time of charge and discharge process were decreased with the increase in the current density. By calculating capacitance, the capacitance of NAHC1 was decreased with the increase in current density in Figure 6c. At a low current density, decreasing the capacitance was not obvious. However, it badly decreased at a high current density. Capacitance at 1 A/g was 225.4 F/g, while capacitance at 10 A/g was 151 F/g. The capacitance of NAHC1 at 10 A/g remained at 67%. Table 3 shows the performance comparison of NAHC1 with other biomass carbon materials.

4. Conclusions

In this study, different kinds of carbon material from bio-oil were fabricated through hydrothermal pretreatment, nitrogen doping, and KOH activation. The NHCs had a honeycomb-like structure, which promoted the transmission of electrolytes. However, the NHCs had a poor pore structure and a specific surface area of just 14.3789 m2/g; thus, the specific surface of NAHCs was enhanced greatly after KOH activation, which had a specific surface area of 1163.3115 m2/g. The NAHCs had a lower graphitization degree and ID/IG than the NHCs. The nitrogen content was improved by nitrogen doping. NAHC1 had a nitrogen content of 3.47%. The nitrogen doping types of NAHC1 were mainly pyridinic-N(N-6), pyrrolic-N(N-5), and quaternary-N(N-Q), which contribute to the enhancement of the electrochemical performance. Compared with other NAHCs, NAHC1 had a higher nitrogen content. Thus, NAHC1 had the best electrochemical performance of 225.4 F/g at 1 A/g. Compared with other NAHCs, NAHC1 had nitrogen content. Then NAHC2 had an electrochemical performance of 183.3 F/g and NAHC3 had an electrochemical performance of 175.3 F/g at 1 A/g. As a result of poor structure, NHCs had a very low electrochemical performance of 63.2–82.8 F/g at 1 A/g. In conclusion, carbon material electrochemical performance factors are varied. In this study, the optimal synthesis sample was NAHC1.

Author Contributions

Conceptualization, J.W. and H.Z.; Data curation, J.W., X.S. and B.L.; Funding acquisition, D.X.; Investigation, J.W., J.S., S.Z., M.W., L.S. and B.L.; Methodology, J.W. and S.Z.; Supervision, D.X., S.Z. and H.Z.; Writing—original draft, J.S.; Writing—review and editing, J.W., D.X., S.Z. and X.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work is financially supported by National Natural Science Foundation of China (22208164, 32271789), Natural Science Foundation of the Jiangsu Higher Education Institutions of China (22KJB470017), and State Key Laboratory of High efficiency Utilization of Coal and Green Chemical Engineering, Ningxia University (2022-K39).

Institutional Review Board Statement

This study does not involve any ethical issues.

Informed Consent Statement

Not applicable.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gou, G.J.; Huang, F.; Jiang, M.; Li, J.Y.; Zhou, Z.W. Hierarchical porous carbon electrode materials for supercapacitor developed from wheat straw cellulosic foam. Renew. Energy 2020, 149, 208–216. [Google Scholar] [CrossRef]
  2. Taer, E.; Febriyanti, F.; Mustika, W.S.; Taslim, R.; Agustino, A.; Apriwandi, A. Enhancing the performance of supercapacitor electrode from chemical activation of carbon nanofibers derivedAreca catechuhusk via one-stage integrated pyrolysis. Carbon Lett. 2021, 31, 601–612. [Google Scholar] [CrossRef]
  3. Faraji, S.; Ani, F.N. The development supercapacitor from activated carbon by electroless plating-A review. Renew. Sustain. Energy Rev. 2015, 42, 823–834. [Google Scholar] [CrossRef]
  4. Tie, D.; Huang, S.F.; Wang, J.; Ma, J.M.; Zhang, J.J.; Zhao, Y.F. Hybrid energy storage devices: Advanced electrode materials and matching principles. Energy Storage Mater. 2019, 21, 22–40. [Google Scholar] [CrossRef]
  5. Motoc, S.; Manea, F.; Baciu, A.; Orha, C.; Pop, A. Electrochemical Method for Ease Determination of Sodium Diclofenac Trace Levels in Water Using Graphene-Multi-Walled Carbon Nanotubes Paste Electrode. Int. J. Environ. Res. Public Health 2022, 19, 29. [Google Scholar] [CrossRef] [PubMed]
  6. Islam, M.A.; Ong, H.L.; Villagracia, A.; Halim, K.A.A.; Ganganboina, A.B.; Doong, R.E.A. Biomass-derived cellulose nanofibrils membrane from rice straw as sustainable separator for high performance supercapacitor. Ind. Crops Prod. 2021, 170, 113694. [Google Scholar] [CrossRef]
  7. Zhu, Y.; Fang, T.T.; Hua, J.Q.; Qiu, S.J.; Chu, H.L.; Zou, Y.J.; Xiang, C.L.; Huang, P.R.; Zhang, K.X.; Lin, X.C.; et al. Biomass-Derived Porous Carbon Prepared from Egg White for High-performance Supercapacitor Electrode Materials. Chemistryselect 2019, 4, 7358–7365. [Google Scholar] [CrossRef]
  8. Song, M.Y.; Zhou, Y.H.; Ren, X.; Wan, J.F.; Du, Y.Y.; Wu, G.; Ma, F.W. Biowaste-based porous carbon for supercapacitor: The influence of preparation processes on structure and performance. J. Colloid Interface Sci. 2019, 535, 276–286. [Google Scholar] [CrossRef] [PubMed]
  9. Du, J.; Zhang, Y.; Lv, H.J.; Hou, S.L.; Chen, A.B. Yeasts-derived nitrogen-doped porous carbon microcapsule prepared by silica-confined activation for supercapacitor. J. Colloid Interface Sci. 2021, 601, 467–473. [Google Scholar] [CrossRef] [PubMed]
  10. Bassey, E.; Yang, L.; Cao, M.J.; Feng, Y.; Yao, J.F. Molten salt synthesis of capacitive porous carbon from Allium cepa (onion) for supercapacitor application. J. Electroanal. Chem. 2021, 881, 114972. [Google Scholar] [CrossRef]
  11. Hamouda, H.A.; Cui, S.Z.; Dai, X.W.; Xiao, L.L.; Xie, X.; Peng, H.; Ma, G.F. Synthesis of porous carbon material based on biomass derived from hibiscus sabdariffa fruits as active electrodes for high-performance symmetric supercapacitors. Rsc Adv. 2021, 11, 354–363. [Google Scholar] [CrossRef] [PubMed]
  12. Wang, B.L.; Li, L.; Zhao, C.H.; Zhao, C.J. Microporous active carbon with ultrahigh surface area from Metaplexis japonica for high-performance supercapacitor. Diam. Relat. Mater. 2021, 118, 108484. [Google Scholar] [CrossRef]
  13. Wang, Y.J.; Zhao, L.C.; Peng, H.; Dai, X.W.; Liu, X.N.; Ma, G.F.; Lei, Z.Q. Three-dimensional honeycomb-like porous carbon derived from tamarisk roots via a green fabrication process for high-performance supercapacitors. Ionics 2019, 25, 4315–4323. [Google Scholar] [CrossRef]
  14. Zou, Z.H.; Lei, Y.; Li, Y.M.; Zhang, Y.H.; Xiao, W. Nitrogen-Doped Hierarchical Meso/Microporous Carbon from Bamboo Fungus for Symmetric Supercapacitor Applications. Molecules 2019, 24, 3677. [Google Scholar] [CrossRef] [Green Version]
  15. Xia, C.K.; Surendran, S.; Ji, S.; Kim, D.; Chae, Y.J.; Kim, J.; Je, M.; Han, M.K.; Choe, W.S.; Choi, C.H.; et al. A sulfur self-doped multifunctional biochar catalyst for overall water splitting and a supercapacitor from Camellia japonica flowers. Carbon Energy 2022, 4, 491–505. [Google Scholar] [CrossRef]
  16. Devarajan, J.; Arumugam, P. Boron-doped activated carbon from the stems of Prosopis juliflora as an effective electrode material in symmetric supercapacitors. J. Mater. Sci.-Mater. Electron. 2022, 33, 17469–17482. [Google Scholar] [CrossRef]
  17. Gopalakrishnan, A.; Badhulika, S. Effect of self-doped heteroatoms on the performance of biomass-derived carbon for supercapacitor applications. J. Power Sources 2020, 480, 228830. [Google Scholar] [CrossRef]
  18. Gunasekaran, S.S.; Gopalakrishnan, A.; Subashchandrabose, R.; Badhulika, S. Single Step, Direct Pyrolysis Assisted Synthesis of Nitrogen-Doped Porous Carbon Nanosheets Derived from Bamboo wood for High Energy Density Asymmetric Supercapacitor. J. Energy Storage 2021, 42, 103048. [Google Scholar] [CrossRef]
  19. Ma, G.F.; Yang, Q.; Sun, K.J.; Peng, H.; Ran, F.T.; Zhao, X.L.; Lei, Z.Q. Nitrogen-doped porous carbon derived from biomass waste for high-performance supercapacitor. Bioresour. Technol. 2015, 197, 137–142. [Google Scholar] [CrossRef]
  20. He, D.P.; Niu, J.; Dou, M.L.; Ji, J.; Huang, Y.Q.; Wang, F. Nitrogen and oxygen co-doped carbon networks with a mesopore-dominant hierarchical porosity for high energy and power density supercapacitors. Electrochim. Acta 2017, 238, 310–318. [Google Scholar] [CrossRef]
  21. Wu, L.; Cai, Y.M.; Wang, S.Z.; Li, Z.Y. Doping of nitrogen into biomass-derived porous carbon with large surface area using N-2 non-thermal plasma technique for high-performance supercapacitor. Int. J. Hydrog. Energy 2021, 46, 2432–2444. [Google Scholar] [CrossRef]
  22. Li, J.T.; Xiao, R.; Li, M.; Zhang, H.Y.; Wu, S.L.; Xia, C.L. Template-synthesized hierarchical porous carbons from bio-oil with high performance for supercapacitor electrodes. Fuel Process. Technol. 2019, 192, 239–249. [Google Scholar] [CrossRef]
  23. Wang, Q.; Qin, B.; Zhang, X.H.; Xie, X.L.; Jin, L.E.; Cao, Q. Synthesis of N-doped carbon nanosheets with controllable porosity derived from bio-oil for high-performance supercapacitors. J. Mater. Chem. A 2018, 6, 19653–19663. [Google Scholar] [CrossRef]
  24. Lyu, L.; Seong, K.; 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]
  25. Lin, H.S.; Zhang, L.J.; Zhang, S.; Li, Q.Y.; Hu, X. Hydrothermal carbonization of cellulose in aqueous phase of bio-oil: The significant impacts on properties of hydrochar. Fuel 2022, 315, 123132. [Google Scholar] [CrossRef]
  26. Fu, Y.; Ye, J.; Chang, J.; Lou, H.; Zheng, X.W. Solid fuel production by hydrothermal carbonization of water-like phase of bio-oil. Fuel 2016, 180, 591–596. [Google Scholar] [CrossRef]
  27. Jiang, W.C.; Li, L.Y.; Pan, J.Q.; Senthil, R.A.; Jin, X.; Cai, J.Q.; Wang, J.; Liu, X.G. Hollow-tubular porous carbon derived from cotton with high productivity for enhanced performance supercapacitor. J. Power Sources 2019, 438, 226936. [Google Scholar] [CrossRef]
  28. Wang, Q.; Qin, B.; Li, H.X.; Zhang, X.H.; Tian, X.; Jin, L.E.; Cao, Q. Honeycomb-like carbon with tunable pore size from bio-oil for supercapacitor. Microporous Mesoporous Mater. 2020, 309, 110551. [Google Scholar] [CrossRef]
  29. Cao, L.H.; Li, H.L.; Xu, Z.X.; Zhang, H.J.; Ding, L.H.; Wang, S.Q.; Zhang, G.Y.; Hou, H.Q.; Xu, W.H.; Yang, F.; et al. Comparison of the heteroatoms-doped biomass-derived carbon prepared by one-step nitrogen-containing activator for high performance supercapacitor. Diam. Relat. Mater. 2021, 114, 108316. [Google Scholar] [CrossRef]
  30. Mehare, M.D.; Deshmukh, A.D.; Dhoble, S.J. Preparation of porous agro-waste-derived carbon from onion peel for supercapacitor application. J. Mater. Sci. 2020, 55, 4213–4224. [Google Scholar] [CrossRef]
  31. de Souza, L.K.C.; Martins, J.C.; Oliveira, D.P.; Ferreira, C.S.; Goncalves, A.A.S.; Araujo, R.O.; Chaar, J.D.; Costa, M.J.F.; Sampaio, D.V.; Passos, R.R.; et al. Hierarchical porous carbon derived from acai seed biowaste for supercapacitor electrode materials. J. Mater. Sci.-Mater. Electron. 2020, 31, 12148–12157. [Google Scholar] [CrossRef]
  32. Rajesh, M.; Manikandan, R.; Park, S.; Kim, B.C.; Cho, W.J.; Yu, K.H.; Raj, C.J. Pinecone biomass-derived activated carbon: The potential electrode material for the development of symmetric and asymmetric supercapacitors. Int. J. Energy Res. 2020, 44, 8591–8605. [Google Scholar] [CrossRef]
  33. Song, X.Y.; Ma, X.L.; Li, Y.; Ding, L.; Jiang, R.Y. Tea waste derived microporous active carbon with enhanced double-layer supercapacitor behaviors. Appl. Surf. Sci. 2019, 487, 189–197. [Google Scholar] [CrossRef]
  34. Zhang, R.J.; Zhang, W.; Shi, M.P.; Li, H.; Ma, L.N.; Niu, H.J. Morphology controllable synthesis of heteroatoms-doped carbon materials for high-performance flexible supercapacitor. Dye. Pigment. 2022, 199, 109968. [Google Scholar] [CrossRef]
  35. Yan, D.; Liu, L.; Wang, X.Y.; Xu, K.; Zhong, J.H. Biomass-Derived Activated Carbon Nanoarchitectonics with Hibiscus Flowers for High-Performance Supercapacitor Electrode Applications. Chem. Eng. Technol. 2022, 45, 649–657. [Google Scholar] [CrossRef]
  36. Li, Q.; Lu, T.Y.; Wang, L.L.; Pang, R.X.; Shao, J.W.; Liu, L.L.; Hu, X. Biomass based N-doped porous carbons as efficient CO2 adsorbents and high-performance supercapacitor electrodes. Sep. Purif. Technol. 2021, 275, 119204. [Google Scholar] [CrossRef]
  37. Zhang, Q.T.; Dai, Q.Q.; Yan, C.; Su, C.; Li, A. Nitrogen-doped porous carbon nanoparticle derived from nitrogen containing conjugated microporous polymer as high performance lithium battery anode. J. Alloys Compd. 2017, 714, 204–212. [Google Scholar] [CrossRef]
  38. Song, Y.; Qu, W.W.; He, Y.H.; Yang, H.X.; Du, M.; Wang, A.J.; Yang, Q.; Chen, Y.Q. Synthesis and processing optimization of N-doped hierarchical porous carbon derived from corncob for high performance supercapacitors. J. Energy Storage 2020, 32, 101877. [Google Scholar] [CrossRef]
  39. Wang, H.H.; Liu, J.H.; Zhang, K.; Peng, H.R.; Li, G.C. Meso/microporous nitrogen-containing carbon nanofibers with enhanced electrochemical capacitance performances. Synth. Met. 2015, 203, 149–155. [Google Scholar] [CrossRef]
  40. Wang, Y.Y.; Lu, C.S.; Cao, X.F.; Wang, Q.; Yang, G.H.; Chen, J.C. Porous Carbon Spheres Derived from Hemicelluloses for Supercapacitor Application. Int. J. Mol. Sci. 2022, 23, 7101. [Google Scholar] [CrossRef]
  41. Li, X.P.; Liang, P.; Zhang, J.G.; Liu, B. Electrochemical performance of Pleurotus ostreatus-derived carbon prepared by different methods as electrode in supercapacitor. Mater. Chem. Phys. 2022, 278, 125623. [Google Scholar] [CrossRef]
  42. He, X.J.; Zhang, N.; Shao, X.L.; Wu, M.B.; Yu, M.X.; Qiu, J.S. A layered-template-nanospace-confinement strategy for production of corrugated graphene nanosheets from petroleum pitch for supercapacitors. Chem. Eng. J. 2016, 297, 121–127. [Google Scholar] [CrossRef]
  43. Morishita, T.; Soneda, Y.; Tsumura, T.; Inagaki, M. Preparation of porous carbons from thermoplastic precursors and their performance for electric double layer capacitors. Carbon 2006, 44, 2360–2367. [Google Scholar] [CrossRef]
  44. Shen, F.; Su, J.L.; Zhu, L.F.; Qi, X.H.; Zhang, X. Comprehensive utilization of dairy manure to produce glucose and hierarchical porous carbon for supercapacitors. Cellulose 2017, 24, 2571–2579. [Google Scholar] [CrossRef]
  45. Yan, R.; Wang, K.; Tian, X.D.; Li, X.; Yang, T.; Xu, X.T.; He, Y.T.; Lei, S.W.; Song, Y. Heteroatoms in situ-doped hierarchical porous hollow-activated carbons for high-performance supercapacitor. Carbon Lett. 2020, 30, 331–344. [Google Scholar] [CrossRef]
  46. Serafin, J.; Baca, M.; Biegun, M.; Mijowska, E.; Kalenczuk, R.J.; Srenscek-Nazzar, J.; Michalkiewicz, B. Direct conversion of biomass to nanoporous activated biocarbons for high CO2 adsorption and supercapacitor applications. Appl. Surf. Sci. 2019, 497, 143722. [Google Scholar] [CrossRef]
  47. Zhang, Z.J.; He, J.J.; Tang, X.C.; Wang, Y.L.; Yang, B.B.; Wang, K.J.; Zhang, D.Y. Supercapacitors based on a nitrogen doped hierarchical porous carbon fabricated by self-activation of biomass: Excellent rate capability and cycle stability. Carbon Lett. 2019, 29, 585–594. [Google Scholar] [CrossRef]
Figure 1. SEM image of (a) NHC1, (b) NHC2, (c) NHC3, (d) NAHC1, (e) NAHC2, and (f) NAHC3.
Figure 1. SEM image of (a) NHC1, (b) NHC2, (c) NHC3, (d) NAHC1, (e) NAHC2, and (f) NAHC3.
Ijerph 20 01355 g001
Figure 2. N2 adsorption–desorption isotherms of various samples.
Figure 2. N2 adsorption–desorption isotherms of various samples.
Ijerph 20 01355 g002
Figure 3. (a) XRD pattern of NHCs and NAHCs; (b) Raman spectra of NHCs and NAHCs.
Figure 3. (a) XRD pattern of NHCs and NAHCs; (b) Raman spectra of NHCs and NAHCs.
Ijerph 20 01355 g003
Figure 4. (a) C 1s core level of XPS Spectrum of NHC1; (b) N 1s core level of XPS Spectrum of NHC1; (c) C 1s core level of XPS Spectrum of NAHC1; (d) N 1s core level of XPS Spectrum of NAHC1.
Figure 4. (a) C 1s core level of XPS Spectrum of NHC1; (b) N 1s core level of XPS Spectrum of NHC1; (c) C 1s core level of XPS Spectrum of NAHC1; (d) N 1s core level of XPS Spectrum of NAHC1.
Ijerph 20 01355 g004
Figure 5. CV curves of tested samples at (a) 10 mV/s and (b) 100 mV/s; GCD curves of tested sample at (c) 1 A/g and (d) 10 A/g; (e) Nyquist plots of tested samples; (f) the partial enlarged view of the Nyquist plots.
Figure 5. CV curves of tested samples at (a) 10 mV/s and (b) 100 mV/s; GCD curves of tested sample at (c) 1 A/g and (d) 10 A/g; (e) Nyquist plots of tested samples; (f) the partial enlarged view of the Nyquist plots.
Ijerph 20 01355 g005
Figure 6. (a) CV curves, (b) GCD curves, and (c) the capacitance of NAHC1.
Figure 6. (a) CV curves, (b) GCD curves, and (c) the capacitance of NAHC1.
Ijerph 20 01355 g006
Table 1. Porosity parameters of tested samples.
Table 1. Porosity parameters of tested samples.
SamplePore Structure Parameter
SBet (m2/g)Smcro (m2/g)Smeso (m2/g)Vmcro (m3/g)Vmeso (m3/g)
NHC115.133.8411.290.040.03
NHC214.382.9911.390.010.06
NHC316.53016.5300.08
NAHC11175.46946.18229.280.440.43
NAHC21420.171319.06101.110.550.29
NAHC31163.311079.8283.500.440.24
Table 2. Elemental contents of tested samples measured by XPS.
Table 2. Elemental contents of tested samples measured by XPS.
SampleElemental Composition
C (At.%)N (At.%)O (At.%)
NHC181.773.5514.69
NHC282.193.6314.18
NHC382.063.7814.16
NAHC191.682.935.39
NAHC290.802.846.36
NAHC390.013.306.69
Table 3. Various precursors based on activated carbon as supercapacitor electrodes.
Table 3. Various precursors based on activated carbon as supercapacitor electrodes.
Carbon PrecursorCg (F/g)ElectrolyteRef.
Coal tar pitch1436 M KOH[42]
Poly(vinyl alcohol)2106 M KOH[43]
Dairy manure2026 M KOH[44]
Setaria viridis1996 M KOH[45]
Lumpy bracket2236 M KOH[46]
Glebionis coronaria2056 M KOH[47]
Bio-oil225.46 M KOHThis work
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

Wei, J.; Sun, J.; Xu, D.; Shi, L.; Wang, M.; Li, B.; Song, X.; Zhang, S.; Zhang, H. Preparation and Electrochemical Performance of Bio-Oil-Derived Hydrochar as a Supercapacitor Electrode Material. Int. J. Environ. Res. Public Health 2023, 20, 1355. https://doi.org/10.3390/ijerph20021355

AMA Style

Wei J, Sun J, Xu D, Shi L, Wang M, Li B, Song X, Zhang S, Zhang H. Preparation and Electrochemical Performance of Bio-Oil-Derived Hydrochar as a Supercapacitor Electrode Material. International Journal of Environmental Research and Public Health. 2023; 20(2):1355. https://doi.org/10.3390/ijerph20021355

Chicago/Turabian Style

Wei, Juntao, Jiawei Sun, Deliang Xu, Lei Shi, Miao Wang, Bin Li, Xudong Song, Shu Zhang, and Hong Zhang. 2023. "Preparation and Electrochemical Performance of Bio-Oil-Derived Hydrochar as a Supercapacitor Electrode Material" International Journal of Environmental Research and Public Health 20, no. 2: 1355. https://doi.org/10.3390/ijerph20021355

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