You are on page 1of 36

Energy Technology

Generation, Conversion, Storage, Distribution

Accepted Article
Title: Self-nitrogen-doped porous biocarbon from watermelon rind: a high-
performance supercapacitor electrode and its improved electrochemical
performance with redox additive electrolyte

Authors: Qiu-Feng Lu, Xiao-Qiang Lin, Rui Liu, and Ning Yang

This manuscript has been accepted after peer review and appears as an
Accepted Article online prior to editing, proofing, and formal publication of the
final Version of Record (VoR). This work is currently citable by using the Digital
Object Identifier (DOI) given below. The VoR will be published online in Early
View as soon as possible and may be different to this Accepted Article as a
result of editing. Readers should obtain the VoR from the journal website shown
below when it is published to ensure accuracy of information. The authors are
responsible for the content of this Accepted Article.

To be cited as: Energy Technol. 10.1002/ente.201800628

Link to VoR: http://dx.doi.org/10.1002/ente.201800628

www.entechnol.de
Energy Technology 10.1002/ente.201800628

Self-nitrogen-doped porous biocarbon from watermelon rind: a high-


performance supercapacitor electrode and its improved
electrochemical performance by using redox additive electrolyte

Xiao-Qiang Lin,[a] Ning Yang,[a] Qiu-Feng Lü, *[a] Rui Liu[b]

[a]
Key Laboratory of Eco-materials Advanced Technology (Fuzhou University), College of Materials

Accepted Manuscript
Science and Engineering, Fuzhou University, Fuzhou 350116, China

[b]
Ministry of Education Key Laboratory of Advanced Civil Engineering Material, School of

Materials Science and Engineering, and Institute for Advanced Study, Tongji University, Shanghai,

201804, China

Abstract

Self-nitrogen-doped biocarbons were prepared via an easy one-step pyrolysis-activation process

from watermelon rind. Porous biocarbon ACW-2, obtained with a watermelon/potassium hydroxide

mass ratio of 1:2 at 700 °C, exhibited a large specific surface area of 1303.3 m2 g-1, and was

decorated with abundant nitrogen/oxygen-containing functional groups. When ACW-2 was used as a

supercapacitor electrode, a high specific capacitance of 260 F g-1 was achieved at a current density

of 1 A g-1. ACW-2 also exhibited an ultrahigh cycling stability, and its retention rate was maintained

at 93.8% even after 5000 cycles at a scan rate of 100 mV s-1. Furthermore, ACW-2 acquired a

supramaximal specific capacitance (852 F g-1) and a high energy density (12.9 Wh kg-1) when p-

                                                             
*
Corresponding author:
Tel.: +86 591 22 866 540; Fax: +86 591 22 866 539.
E-mail addresses: qiufenglv@163.com (Q.-F. Lü).

 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

phenylenediamine redox additive was added into the potassium hydroxide electrolyte. Therefore,

watermelon rind can be used as a promising biomass precursor, and the as-prepared porous

biocarbon is expected to be an outstanding comprehensive electrode material for electrochemical

supercapacitors.

Keywords:Nitrogen doping; porous biocarbon; pyrolysis; supercapacitor

Accepted Manuscript
1. Introduction

In recent years, porous carbon materials have been receiving a significant amount of attention

because of their extensive use in soil remediation,[1] adsorption of heavy metal ions,[2] and

supercapacitor electrodes,[3] owing to their large surface areas, appropriate hierarchical porous

structures, and excellent electrical conductivity. Furthermore, the surfaces of porous carbons can be

embellished with various active functional groups to obtain high-value-added products, thereby

broadening the application fields of porous carbons.[4] Biomass-based porous carbons with abundant

pore structures and oxygen-containing functional groups are derived from the high-temperature

anaerobic pyrolysis of organic biomass materials.[5] In porous biocarbons, in addition to the C

element, N, P, and K contents resulting from the residues of amino acids and proteins after high-

temperature pyrolysis are relatively high. Both large surface areas and the doping of N and O

elements into C frameworks can improve the specific capacitance of porous biocarbons;[5,6]

therefore, porous biocarbons tend to be used as efficient supercapacitor electrodes.

Supercapacitors, also known as ultracapacitors, can be classified into two categories according

to the working mechanism, namely the electric double-layer capacitor and the pseudocapacitor.

Supercapacitors have gained traction in research owing to their outstanding characteristics of energy

conversion and storage.[7,8] Up to now, transition metal oxides, carbon materials, conductive


 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

polymers, and newly metal–organic frameworks, have been extensively used as supercapacitor

electrodes. [9-12]

Various carbon materials that include carbon foams,[8] fullerene,[13] nanotubes,[14] carbon

nanohorns,[15] graphene,[16] and microporous carbons,[12] are employed as supercapacitor electrodes

because of their superior performance. However, their commercial use is not widespread due to the

limitations imposed by complicated production processes and high costs. Porous biocarbons have

Accepted Manuscript
several exceptional advantages such as simple fabrication method, low cost, renewability, and wide

availability. Especially, nitrogen-doped porous biocarbons are not only effective in increasing the

wettability on electrode and electrolyte surfaces, but can also generate pseudocapacitance.[17]

Moreover, hierarchical porous structures, namely the micropores (﹤2 nm) that can store charge

during the charging–discharging process and ensure large specific surface areas, the mesopores (2–

50 nm) that can provide transfer channels for electrolyte ions, and macropores (﹥50 nm) that serve

as repositories for electrolytes facilitating the contact of electrolyte and electrodes,[17, 18] greatly

boost the specific capacitance of the supercapacitor.

To date, plenty of biomass materials, such as pomelo peel,[6] sugarcane bagasse,[19] shiitake

mushroom,[20] willow catkin,[21] kapok fiber,[7] eggplant,[22] banana peel,[23] and corncob residue,[24]

have been used to prepare porous biocarbons into excellent supercapacitor electrodes through

carbonization and activation, as has been reported in the literature.

As a biomass waste, watermelon rind has not been fully utilized and is being disposed to the

environment in large amounts every year. Up to now, only a few studies have reported that

watermelon rind may be regarded as a raw material in the preparation of high-performance porous

carbons.[25, 26] Environmental pollution can also be caused when it decomposes naturally, and not

just when it is disposed as biological resource waste. Watermelon rind comes from the outer skin of

 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

watermelons and contains waxiness, citrulline, and polysaccharides (fructose and glucose). The

porous carbon skeleton can be left behind after polysaccharides decompose into many small

molecules through anaerobic pyrolysis, while the nitrogen heteroatoms in the carbon structures

originate from citrulline.[27, 28] Therefore, watermelon rind can be used as an excellent precursor for

porous biocarbon electrodes and this is a direct and efficient approach to solving the problem of

biomass waste.

Accepted Manuscript
As we know, introducing faradaic reaction into supercapacitors can tremendously improve

their specific capacitance. Electrolytes are significant and act as an ion transport medium during the

charge–discharge process in the supercapacitor. Therefore, using a redox mediator added to the

electrolyte is an innovative and low-cost method to achieve excellent electrochemical

supercapacitor performance. Thus far, VOSO4,[29] KI,[30] methylene blue,[31] p-phenylenediamine

(PPD),[32] etc. have been thought as high-efficiency redox additives. Among these, PPD is an

excellent choice because it can simultaneously produce two protons and two charges due to the

conversion of intrinsic redox states through redox reaction. Thus, faradaic reaction is introduced

and the specific capacitances of the supercapacitor are increased.[33]

In this study, we adopted an eco-friendly and cost-efficient strategy to construct self-nitrogen-

doped porous biocarbons by using watermelon rind as a carbon precursor through one-step

potassium hydroxide (KOH) activation without precarbonization. As a result, biocarbon materials

with high specific surface areas and adequate porosities were achieved. The self-nitrogen-doping of

the biocarbon originated from the citrulline in the watermelon rind during the high-temperature

pyrolysis process, and then from increasing the wettability and pseudocapacitance of the biocarbon.

Subsequently, a redox additive PPD was added into the electrolyte and it was expected that the

capacitance and the resultant energy density of the biocarbon would be markedly improved by the

 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

redox reaction that occurred at the positive and negative electrodes. Due to its high surface area,

high conductivity, and efficient porous ion transport pathways, the as-obtained biocarbon showed

great specific capacitance and energy density with super-high cycling stability.

2. Experimental section

2.1. Materials

Accepted Manuscript
Fresh watermelon rind was acquired from a local supermarket, cut into small pieces, and

baked in the sun for five days. Subsequently, the dry rind was made into powder by a pulverizer.

Potassium hydroxide (KOH) and PPD were obtained from Sinopharm Chemical Reagent Co., Ltd.,

and the Shanghai Chemical Reagent Company (Shanghai, China), respectively and were used

directly. Hydrochloric acid (HCl) was received from Lanxi Xuri Chemical Engineering Co., Ltd.

(Zhejiang, China) and diluted to a concentration of 1 M. The air-laid papers were purchased from

Shenzhen Chenyan Trading Co., Ltd. (Guangdong, China).

2.2. Preparation of the nitrogen-doped porous carbons

Nitrogen-doped porous biocarbons were prepared by an easy one-step pyrolysis-activation

method. The watermelon rind powder (1 g) and appropriate KOH were placed directly into 30 mL

of deionized water and stirred for 5 h. The solution of the mixture was dried at 70 °C for 24 h to

obtain a precursor. Then, the precursors were pyrolyzed at 700 °C with a heating rate of 5 °C min-1

for 1 h under nitrogen flow. Subsequently, the gained intermediate products were washed by 1 M of

HCl solution and were then dried at 70 °C for 24 h. The obtained porous biocarbons (ACW) are

denoted as ACW-1, ACW-2, and ACW-3, and indicate that the mass ratio of KOH to the raw material

is 1:1, 2:1, and 3:1, respectively. The conversion yields of ACW-1, ACW-2, and ACW-3 are 9.84%,

 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

8.65% and 6.32%, respectively. These lower conversion yields are attributed to the one-step

pyrolysis-activation preparation process from low-carbon content watermelon rind raw materials.

The typical procedure for the preparation of ACW-2 as an electrode is illustrated in Scheme 1. CW-

700 was obtained from the directed carbonization of the watermelon rind at 700 °C without adding

KOH, and its conversion yield was about 33.2%.

Accepted Manuscript
2.3. Characterization

Scanning electron microscopy (SEM, Carl Zeiss ULTRA 55) was used to represent the surface

morphology of biocarbon products. Thermogravimetric analysis (TGA) was characterized by using

a Q600 thermal analyzer. The asymmetric constructions of functional groups were characterized by

the FT-IR spectra through a Nicolet FT-IR 5700 spectrophotometer. Phase information was

performed on X-ray diffractometer (XRD) patterns by using a Rigaku wide-angle XRD system. A

HORIBA Jobin-Yvon Raman spectrometer was applied to Raman spectra analysis. A Vario EL Cube

elemental analyzer was used for elemental analysis. The X-ray photoelectron spectrum (XPS) was

obtained through an X-ray spectrometer (ESCALAB 250). The Brunauer–Emmett–Teller (BET)

specific surface areas and pore size distributions were observed by using a Micromeritics 3Flex

analyzer.

2.4. Electrochemical measurements

Electrochemical performance of the samples was measured at room temperature by using a

CHI660E electrochemical workstation purchased from Chenhua Instruments Co. The individual

electrode testing device was a typical three-electrode system that included a platinum wire,

Ag/AgCl, and working electrodes. The Ag/AgCl reference electrode was immersed in 3 M KCl


 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

solution. Moreover, two 6 M KOH solutions contained different PPD concentrations as electrolytes

during the testing process. The working electrodes were fabricated as follows: in brief, fully grinded

biocarbon materials, acetylene black conductive additive, and poly(tetrafluoroethylene) binder,

were mixed to obtain slurry with a mass ratio of 85:10:5. The slurry was pressed onto a stainless

steel mesh (1 cm  1 cm) under a pressure of 10 MPa for 1 min, and was finally dried at 70 °C for

12 h. Subsequently, a mass of approximately 4 mg cm-2 was loaded onto a current collector, and the

Accepted Manuscript
thickness of working electrode was about 0.15 mm. The specific capacitances of the biocarbon were

obtained through computation from the cyclic voltammetry (CV) curve areas with a potential

window of -1.0 to 0 V, or by calculation from the discharge time of the galvanostatic charge–

discharge (GCD) tests at a current density range of 1–10 A g-1. Under the scanning frequencies in

the 0.01–100 kHz range, the resistance information of the charge and ions were explored through

electrochemical impedance spectroscopy.

The symmetrical supercapacitor was assembled by the sandwich method by using an air-laid

paper as a separator inserted into the two identical electrodes. The values of specific capacitance (C,

F g-1), energy density (E, Wh kg-1), and power density (P, W kg-1) were obtained by using the

following equations:

C= (It) / (mV) (1)

E= (CV2) / (23.6) (2)

P= (E3600) / t (3)

where I (A) is the discharge current, t (s) represents discharge time, m is the total mass of

biocarbon materials in the two electrodes, and V (V) is the potential window, and V (V) is the

cell-operation potential.


 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

3. Results and discussion

3.1. Characterization of the nitrogen-doped porous biocarbons

The SEM image of the CW-700 sample in Figure 1a shows a caking and melting texture and

some pores that are not obvious. The crooked and shrinking profile of CW-700 is attributed to the

contraction and dehydration of the cell walls, and the pores are formed from the volatilization and

Accepted Manuscript
deprivation of small molecules in the watermelon rind during the pyrolysis process.[34] In

comparison to CW-700, the connected porous network structures of the biocarbon samples (Figure

1b, 1c, 1d) increase significantly. This is because KOH is a pore-foaming agent that can etch carbon

materials to bring up larger specific areas and hierarchical pores.[35] Although the biocarbon samples

possess some macropores, as shown in Figure 1, in relation to ACW-1 and ACW-3, ACW-2 have

more abundant and smaller pore construction. This is credited to the added amount of KOH lacking

in ACW-1, since sufficient etching could not be generated. On the other hand, the amount of the

pore-foaming agent is considerable in ACW-3 and leads to excessive etching and abundant

combined macropores that result from the collapse of micropores and mesopores.[36] The ACW-2

biocarbon possesses more intensive small pores, suggesting that it had a large specific area and

excellent electrochemical behavior.

The pyrolysis behavior of the watermelon rind could be assessed intuitively from the TGA–

DTG curves in Figure 2a. According to the DTG curve, the pyrolysis process was roughly divided

into five steps. The first step appeared at 50–100 °C, and was related to the evaporation of moisture

in the watermelon rind. When the temperature interval was 150–200 °C, there was significant mass

loss that was consistent with the gaseous products derived from the decomposition of bound water

and the initial pyrolysis of polysaccharides.[37] As the temperature increased, an intensive peak of

 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

mass loss occurred at approximately 300 °C due to the deep thermolysis of polysaccharides

(fructose and glucose). [38] Two feeblish mass losses occurred at approximately 400 °C and 1,000 °C,

and were credited to the pyrogenic decomposition of citrulline and heteroatoms, respectively. When

the pyrolysis temperature reached 700 °C, 32% of the original mass from the TGA curve still

remained. Therefore, 700 °C was the optimal carbonization temperature at which the watermelon

rind could be pyrolyzed more effectively. Furthermore, the as-prepared biocarbon could maintain a

Accepted Manuscript
mass of heteroatoms, which were beneficial to the electrochemical performance of the

supercapacitor.

The FT-IR spectra of the biocarbon samples are presented in Figure 2b. The broad and feeble

peak of the biocarbon materials at approximately 3,400 cm-1 may have been due to the stretching

vibrations of O-H and N-H.[39] The two characteristic peaks were located at 1,200 cm-1 and 1,350

cm-1, and corresponded to C-O and C-N stretching vibrations, respectively.[40] The oxygen- and

nitrogen-containing groups that existed in the biocarbon samples may have been due to the

polysaccharides and citrulline in the watermelon rind after annealing. Another obvious adsorption

peak at approximately 1,590 cm-1 was ascribed to the C=C tensile deformation vibration of benzene

rings, which suggested that a part of carbons had been graphitized.[41] It was concluded that the

nitrogen atoms from the raw material had been doped into carbon construction and some

graphitized carbon appeared in the biocarbon samples.

Figure 2c exhibits CW-700, ACW-1, ACW-2, and ACW-3 XRD curves and there are two broad

peaks centered at approximately 24° and 43° in the biocarbons, and belong to the diffraction of the

(002) and (100) lattice planes of graphite crystal, respectively.[42] This implied the presence of

graphitized and amorphous carbons in the biocarbon samples.[43] Distinctly, the characteristic (002)

peak of ACW-2 moved to a lower angle, indicating that the interplanar spacing of ACW-2 was larger

 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

than the other biocarbon samples, and this may have been due to the existence of vast micropores

and a random combination of graphitic and chaotic stacking.[44] Furthermore, the peak at 43° was

very faint, and showed that the extent of graphite development was mild.[45] Moreover, the baseline

at a low ACW-2 angle was raised, and could be credited to the large amount of micropores present in

the carbon structure. [46]

The Raman spectra of biocarbons are shown in Figure 2d, which illustrates that the D band and

Accepted Manuscript
G bands are distinct and located at approximately 1,350 cm-1 and 1,590 cm-1, respectively. As we

know, the defect denominators and the disorder of carbon materials were reflected by the D band,

while the G band had a close correlation with the sp2 hybrid carbon atoms in the graphite. The

intensities of the comparative values (ID/IG) of the D and G bands are usually employed to evaluate

the degree of defects in carbon materials.[47] The ID/IG values of CW-700, ACW-1, ACW-2, and ACW-

3, were 1.02, 1.02, 1.04, and 1.01, respectively, and displayed the generation of long-range graphite

and abundant structural defects.[48] Apparently, ACW-2 had the largest ID/IG ratio among the four

samples, and could reveal the generation of nitrogen-doping and rich micropores, in consistence

with the SEM and XRD results.

Elemental analysis was applied to ascertain the elementary composition of biocarbon samples.

As shown in Table 1, the carbon contents of the ACW samples were higher than those in CW-700.

However, there was a crosscurrent for the heteroatoms. This trend was credited to the heteroatoms

resulting from the incomplete carbonization of the watermelon rind and after KOH activation and

subsequent treatments, a part of the heteroatoms being washed or exposed to air.[49]

Table 1. Elemental analyses of the biocarbon samples.

Samples N (%) C (%) H (%) O (%)

10 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

CW-700 3.41 54.72 3.42 32.24

ACW-1 2.33 52.16 3.20 24.43

ACW-2 2.19 56.35 4.03 28.96

ACW-3 2.59 62.16 3.19 24.92

Additionally, among the ACW samples, ACW-2 possessed the lowest nitrogen content and the

highest oxygen content, and its heteroatomic content was up to 31.35%. This could be attributed to

the increase of the oxygen-containing ACW-2 groups by a moderate KOH corrosion. All the ACW

Accepted Manuscript
samples had high carbon contents over 50%, while the heteroatomic contents exceeded 25%. For

the supercapacitor electrode, high carbon content could effectively improve the cycle stability and

conductivity, while the abundant heteroatoms favored the wettability of electrodes to the electrolyte;

therefore, a high carbon content enhanced the electrochemical activity. [50]

The chemical composition of the ACW-2 surface was certified by elemental binding energy analyses

from XPS, and the results are presented in Figure 3. Three salient peaks, namely, C 1s (285 eV), N

1s (400 eV), and O 1s (532 eV), were displayed, demonstrating the coexistence of C, N, and O

elements in ACW-2 (Figure 3a). With regard to the C 1s spectrum of ACW-2 (Figure 3b), six

branches of peaks constituted the best fitting. One strong intensity peak of C 1s appeared at 284.5

eV, which corresponded to a C-C bond and suggested that the vast sp3 carbon rested in ACW-2.

Furthermore, the existence of the C=C signal (285.3 eV) further confirmed that a portion of

amorphous carbons was transformed into a graphite state. Three peaks of C 1s located at 286.2,

288.0, and 289.0 eV, corresponded to C-O, C=O, and HO-C=O bonds, respectively.[51] This

demonstrated that ACW-2 owned massive oxygen-containing groups after activation. The three

peaks that included C-O (532.8 eV), C=O (531.9 eV), and HO-C=O (533.4 eV) (Figure 3d) also

verified the oxygen-containing group residue in ACW-2. The other peak at 285.2 eV corresponded to

11 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

the C-N bond, proving that the N atoms had been doped into the carbon frameworks. This could be

further confirmed by N 1s from Figure 3c. The graphitic-N peak (401.5 eV), the pyrrolic-N peak

(399.5 eV), and the pyridinic-N peak (398.2 eV) were rooted in the pyrolyzation process of

citrulline in the watermelon rind.

It is well known that graphitic-N is beneficial to the electrical conductivity, whereas pyrrolic-N

and pyridinic-N can improve the wettability and capacitance of the biocarbons. This is because

Accepted Manuscript
nitrogen-containing functional groups possess superior electron-donor, high charge mobility, high

surface energy as well as high pseudocapacitance via additional faradaic redox reactions, enhancing

electrochemical capacity. The faradaic reactions of nitrogen-containing functional groups in KOH

electrolyte are as follows:[ 52]

The contents of pyrrolic-N and pyridinic-N progressively increased in biocarbon samples with

the increase in watermelon/KOH mass ratio. However, the graphitic-N exhibited opposite trends

(Figure 3e). Among these samples, ACW-2 possessed the highest pyrrolic-N and pyridinic-N

contents (Table 2).

Furthermore, in the oxygen-containing functional groups of electrode materials, the –OH, C-O

and C=O are beneficial to the wettability and capacitance of electrodes, respectively. With the

increase in watermelon/KOH mass ratio, the contents of –OH  progressively increased in biocarbon

samples, meaning the electrodes have better wettability (Figure 3f). And the content of C-O and

C=O of ACW-2 reach up to 83.01% (Table 2) with highest oxygen content (Table 1). The oxygen-

containing functional groups in KOH electrolyte can also generate faradaic reaction, and can be

summarized as follows:
12 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

Table 2. Contents of different forms of nitrogen and oxygen in the biocarbon samples obtained from

XPS analysis.
Samples Graphitic-N Pyrrolic-N Pyridinic-N -OH C-O C=O

Accepted Manuscript
(%) (%) (%) (%) (%) (%)
ACW-1 57.38 33.68 8.94 14.85 65.86 21.29
ACW-2 18.72 62.81 18.47 16.99 69.77 13.24
ACW-3 21.45 61.25 17.30 22.52 48.56 28.92

From the above analysis, we can conclude that ACW-2 was decorated with abundant oxygen-

and nitrogen-containing groups, and it can obtain outstanding wettability and capacitance.

To further explore the porous textural structures of the obtained carbons, the N2 adsorption–

desorption isothermals were analyzed in detail. The curves of the samples in Figure 4a apparently

exhibit combined peculiarities of type-I and type-IV isotherms that imply the coexistence of

hierarchical pores. For the ACW samples, there was a fast and obvious adsorption, when the relative

pressure was below 0.1, and a faint hysteresis loop appeared at a high relative pressure over 0.4,

implying that micropores were the main ingredient and that a few mesopores existed in the

carbons[53,54]. Additionally, the small steep adsorption at the relative pressure of 0.9–1.0

demonstrated the existence of macropores.[55] However, CW-700 displayed a typical hysteresis loop,

which was the appearance of the capillary condensation phenomenon and demonstrated the

formation of mesopores.[56] This analysis can be confirmed by the distribution of micropore and

mesopore size (Figure 4b, 4c) and by the pore size distributions of the ACW samples that were

highly concentrated in the micropores. Specifically, except from a few mesopores near the

13 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

micropores, the pore size of ACW-2 was approximately 0.5 nm and the pore sizes of ACW-1 and

ACW-3 were mainly 1 nm. However, the CW-700 pore distribution was not only centered at

approximately 1 nm, but was also within a weak and negligible range of 5–20 nm.

The activated ACW samples possessed higher adsorbed volume than the directly carbonized

CW-700 sample at a small pore size (0.5–1.5 nm). This testified for the production of larger

micropore areas whose characteristics are listed in Table 3. The BET surface area (SBET) of CW-700

Accepted Manuscript
was calculated as 927.7 m2 g-1 and the ratio of the micropore area (Smic/SBET) was only 68.0%.

However, the SBET and Smic/SBET values of ACW-1, ACW-2, and ACW-3 overtly increased to 1,378.5

m2 g-1 and 71.9%, 1,284.2 m2 g-1 and 78.5%, and 1,447.6 m2 g-1 and 76.7%, respectively, signifying

that the micropores were mainly created by the activation process. This was because a mass of

micropores was caused by the etching of the caking carbon texture and the opening of occlusive

pores [57] under the action of high-temperature KOH activation. In comparison to ACW-1 and ACW-3,

the Smic/SBET value of ACW-2 was larger. This trend revealed that a lower dosage of KOH could be

easily inserted into the carbon layers to activate the carbon frameworks for enhancing the micropore

area. However, as the value increased further, excess alkalis started corroding the wall of the

micropores, and caused the broadening of pore size and the collapse of the pore structures, resulting

in the decrease of the Smic/SBET value.[58] From the above analysis, the micropores, mesopores, and

macropores in the obtained samples were confirmed. The micropores could enhance the specific

surface areas, store charge, and optimize the rate of performance; however, for excessively small

micropores, the surface areas could not be adequately applied.[59] A certain amount of mesopores

and macropores provided transfer channels for electrolyte ions and a reservoir for electrolytes; thus,

the specific capacitance of the supercapacitor received a significant boost.

14 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

Table 3. Characteristics of the as-obtained biocarbon porous structures

Samples SBETa Dnb Vtotalc Smicd Smic/SBET

(m2 g-1) (nm) (cm3 g-1) (m2 g-1) (%)

ACW-1 1378.5 0.7-4.0 0.71 991.1 71.9

ACW-2 1284.2 0.4-3.0 0.64 1008.4 78.5

ACW-3 1447.6 0.7-4.0 0.72 1110.8 76.7

CW-700 927.7 0.7-20.0 0.49 630.9 68.0

Accepted Manuscript
a
Specific surface area determined according to BET equation.

b
Pore size.

c
Total pore volume.

d
Micropore surface area from t-plot method.

3.2. Electrochemical performance

To assess the parameters of the as-obtained porous carbons for the supercapacitor electrodes,

their electrochemical performance was measured in a 6 M KOH solution by a three-electrode

system. The CV curves of the four samples at 50 mV s-1 are depicted in Figure 5a. The CV curves

of ACW-1, ACW-2, and ACW-3 had typical nearly rectangular profiles, which proved that the double

layer capacitor behaviors were overwhelming during the charge–discharge process.[24] However,

CW-700 visibly displayed a malformed curvilineal shape that was mainly caused by major

amorphous carbon and a greatly disordered pore structure, and resulted in poor conductivity.[60]

Moreover, there were very small humps for the biocarbon samples at approximately -0.5 V that

were probably due to the development of pseudofaradaic redox reactions originating from the

nitrogen- and oxygen-containing groups on the porous material surfaces.[5] Additionally, the larger

region areas of the closed curves implied higher specific capacitances. Clearly, the CV curve area of
15 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

ACW-2 was the largest among the four samples, and meant that ACW-2 obtained the largest specific

capacitance among the supercapacitor electrodes.

Figure 5b shows a plot of the GCD test results for the four samples at a current density of 1 A

g-1. According to the calculation of discharge time, the specific capacitances of CW-700, ACW-1,

ACW-2, and ACW-3, were 94.3, 205, 260, and 221 F g-1, respectively. Obviously, the specific

capacitances of the ACW activated samples were significantly higher than those of CW-700, because

Accepted Manuscript
of their larger specific areas and abundant micropores. Although the SBET of ACW-2 was lower than

those of ACW-1 and ACW-3, ACW-2 still held the largest specific capacitance. It is possible that

ACW-2 had the highest pyrrolic-N and pyridinic-N contents, a relatively higher Smic/SBET value than

the other samples, [58] and abundant ultramicropores, i.e., < 0.7 nm. [61] The pore size of ACW-2 was

approximately 0.5 nm and the pore sizes of ACW-1 and ACW-3 were approximately 1 nm, as shown

in Figure 4b. According to the study by Chmiola et al.,[61] when the pore size was less than 1 nm in

a porous carbon electrode, the solvation shells became highly distorted to squeeze through the pore.

Such distortion would allow a closer approach of the ion center to the electrode surface, and would

lead to improved capacitance.

The CV curves of ACW-2 at various scan rates within 5–300 mV s-1 are illustrated in Figure 6a.

The nearly rectangular shape is retained well even as the scan rates increased at 300 mV s-1. This

reveals the excellent rate capability of ACW-2 during the fast charge–discharge process that may be

credited to the interworking of micro- and mesoporous structures and surface heteroatoms.[24]

Figure 6b shows the GCD curves of ACW-2 at the current density interval of 1 to 10 A g-1.

Distinctly, the shapes of the GCD curves are symmetrical and closed such that an isosceles triangle

is formed at various current densities. This means that ACW-2 has an ideal capacitive behavior and

good characteristics of capacitive reversibility, when it is used as a supercapacitor electrode.[19] A


16 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

small platform exists at lower current densities and low potentials over -0.1 V due to the electrolyte

ions entering the micropores of ACW-2 more slowly, when the electrodes are under the action of a

small current. Furthermore, there is no obvious IR drop, demonstrating the low ESR of the ACW-2

electrode.[19] At a current density of 1 A g-1, the specific capacitance of ACW-2 reaches up to 260 F

g-1 (Figure 6c), and remains at 184 F g-1 even at the high current density of 10 A g-1. Moreover, the

capacitance retention is 70.8% over the current density range of 1 to 10 A g-1, demonstrating the

Accepted Manuscript
great rate performance of this ACW-2 electrode. At higher current densities, the lower capacitance is

attributed to the narrow pore size of the micropores hindering the effective diffusion of electrolyte

ions, while lessening the accumulation of electrolyte ions on micropore interfaces, leading to the

attenuation of ACW-2 capacitance.[58]

For comparison, the specific capacitance of previously published different activated carbons

from biomass precursors is summarized in Table S1. Moreover, it can be seen that ACW-2 exhibits

outstanding electrochemical performance. Mo et al.[25] also fabricated a nitrogen-doped activated

carbon from watermelon rind and applied it to a supercapacitor electrode; the specific capacitance

reached 333 F g-1, which was higher than that of ACW-2 (260 F g-1). However, they fabricated

nitrogen-doped activated carbon through multiple steps of freeze-drying, hydrothermal treatment,

freeze-drying again, and carbonization/activation. Their complex preparation process and high

equipment requirements imposed a limit to its commercial use. In comparison, we prepared ACW-2

through an easy one-step pyrolysis-activation process. As a supercapacitor electrode, the ACW-2

prepared possesses the advantages of high specific capacitance and a simple synthetic route.

To further investigate the electrochemical information of the obtained biocarbons,

electrochemical impedance spectroscopy was measured to describe the action of electrolyte ions

and the electric charge transfer. Figure 7a shows the Nyquist plots of the porous biocarbon samples.
17 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

The intercept values between the Nyquist plots and the real axis at a high frequency (inset), reflect

the equivalent series resistance (ESR), which represents the sum of the samples’ intrinsic resistances,

ion resistance of electrolytes, and contact resistance of electrodes and current collectors.[17] The

ESR of ACW-2 is only 0.7 Ω, which is smaller than those of other samples. The semicircles detected

in the Nyquist plots conform to the charge-transfer resistance between electrodes and electrolytes.

With regard to ACW-2, the resistance value is minimum (0.6 Ω) due to the ample electron transfer

Accepted Manuscript
pathways supported by micropores.[4, 62] A 45 slope proved the existence of Warburg resistance at

the midfrequency area, and corresponded to the behavior of ions diffusion into the active electrode

materials.[63] The Warburg impedance value of CW-700, at approximately 9 Ω, was much larger than

those of the other three samples. And this maybe attribute to that its caking and stacking structure

without enough pores cannot provide adequate access and increase the difficulty of ions diffusion

into the CW-700 electrode.[24] Furthermore, the ACW-2 gained the smallest Warburg impedance (0.6

Ω) among the four samples. And this probably because hierarchical pores in the ACW-2, especially

ample micropores, offered a great deal of pathways to make electrolyte ions diffuse into electrodes

more easily.[60,64] Additionally, the ideal polarizable capacitance is certified by a vertical bar

parallel to the y-axis at low frequencies. [65] The results presented above illustrate that ACW-2 has

the smallest resistivity and is suitable as a supercapacitor electrode.

Cycling stability plays an important role in evaluating the performance of a supercapacitor.

Figure 7b shows the cycling stabilities of the four biocarbon samples at 100 mV s-1 after 5,000

cycles. It can be seen that the biocarbons still maintained ultrahigh specific capacitance after 5,000

CV cycles, which may have been due to the stable porous biocarbon structures.[3] Especially, the

SEM images of the ACw-2 biocarbon electrode after 5000 cycling test are displayed in Figure 7c

and Figure 7d. Clearly, the framework and porous structure of ACw-2 biocarbon were well
18 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

maintained after the long-term cycling test. This demonstrates that the morphology of the ACw-2

biocarbon electrode has not changed obviously during repeated charge–discharge processes. The

long-term stabilities implies that the biocarbons are excellent electrode materials for supercapacitors.

Moreover, the specific capacitance value of CW-700 was rapidly augmented in the first 700

cycles and increased by 7% even after 5,000 CV cycles, which was probably due to “electro-

activation,” i.e., electrolytes calmly entering into electrodes through wetting action.[14,66] The

Accepted Manuscript
nitrogen doping and porous structures bring about a high ACW-2 stability of 93.8%, due to the

efficient and steady transport of ions and electrons on the electrode surfaces.

To estimate the relationship of PPD and electrochemical performance, the effect of different

PPD concentrations on the specific capacitance of the supercapacitor was investigated. Figure 8a

shows that a pair of intense redox peaks exists at approximately -0.1 V and -0.7 V in the CV curves,

due to the conversion of p-phenylenediamine/p-phenylenediimine on the surfaces between

electrodes and the electrolyte during the redox reaction.[67] With the production of two protons and

the two charges from the redox reaction of PPD in the electrolytic system, the shapes of the CV

curves were subjected to big changes, and their rectangular shape was lost.[33] The addition of PPD

obviously heightened the specific capacitance of ACw-2 by introducing a faradaic reaction[62], which

could be confirmed by the increase in the areas of the CV curves.

The above analysis can be further confirmed by Figure 8b. Compared with the

isosceles triangle profile of the ACw-2 GCD curve that was tested in the KOH electrolyte, there are

two small slopes in the GCD curves during the charging–discharging process in the KOH + PPD

electrolytes. Moreover, when the additional amount of PPD was 1.0 mM, the specific capacitance of

ACw-2 reached up to 852 F g-1 from the discharge time, which was three times as much as that of

ACW-2 only in the KOH electrolyte. Table S2 shows the comparisons of capacitances of electrodes
19 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

in various redox active electrolytes. It clearly illustrated that capacitances of electrodes could be

greatly improved when redox mediators were added into the electrolyte, and PPD was a high-

efficiency redox mediator. During the charging–discharging process, the specific capacitance of

ACw-2 electrode was derived from not only the double-layer capacitance but also faradaic reaction,

and that faradaic reaction played a central role at that point. Furthermore, the high specific areas of

ACw-2 (1,303.3 m2 g-1) increased the contact area of electrode material and electrolyte, providing

Accepted Manuscript
sufficient reaction zone for redox reaction of PPD in the electrode/electrolyte interface. However,

the excess PPD concentration in the KOH+PPD electrolyte caused the phase separation of the

electrolyte and gathered free ions and charge, which decreased the conductivity and capacitive

performance of ACw-2.[68] Thus, by adding 1.0 mM PPD into KOH, the electrolyte was optimal for

the purpose of acquiring the maximum specific capacitance.

To further realize the characteristics of the ACW-2 electrode for supercapacitors, ACW-2//ACW-

2 symmetrical cells were assembled in the 6 M KOH and 6 M KOH + 1 mM PPD electrolytes,

respectively. From Figure 9a, the specific capacitances of the two-electrodes, based on the total

mass of the ACW-2 active materials, are up to 54.6, 47.8, 42.8, and 35.0 F g-1 at current densities of

0.5, 1.0, 2.0, and 5.0 A g-1, respectively. Figure 9b validates that adding 1 mM PPD into 6 M KOH

of electrolyte could overtly improve the specific capacitances of the ACW-2//ACW-2 symmetrical

cells, which increased to 93.1, 59.9, 50.6, and 41 F g-1 at 0.5, 1.0, 2.0, and 5.0 A g-1, respectively,

due to the introduction of faradaic reaction.[33] And the charge-discharge rates at 0.5 A g-1 was 0.54.

The relationship of energy density and power density is embodied in the ragone plots (Figure 9c).

At a power density of 0.25 kW kg-1, the energy density was found to be 7.6 and 12.9 Wh kg-1 in 6

M KOH and 6 M KOH + 1 mM PPD of electrolyte, respectively. Additionally, the energy densities

were still maintained at 4.9 and 5.7 Wh kg-1, even at the higher power density of 2.5 kW kg-1. Table
20 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

S3 shows the comparisons of the energy density and power density of ACW-2 to other various

biocarbons, demonstrating clearly that ACW-2 can be applied as an efficient energy-storing material

for commercial use. In general, ACW-2 could be used as an outstanding supercapacitor electrode.

Conclusions

Self-nitrogen-doped porous biocarbons obtained from watermelon rind exhibit large specific

Accepted Manuscript
surface areas, extraordinary porous structures and superexcellent electrochemical performance. The

doping of N and O heteroatoms introduces pseudocapacitances, which increase the specific

capacitances of biocarbons. ACW-2 acquired a supernal specific capacitance (260 F g-1) and

preeminent long cycle stability (93.8%). A supramaximal specific capacitance (852 F g-1) and high

energy density (12.9 Wh kg-1) were obtained when p-phenylenediamine was added into the KOH

electrolyte. Therefore, an eco-friendly and cost-efficient strategy was developed in the construction

of ACW-2 porous biocarbon, and is expected to be commercially produced as a high-performance

electrode for supercapacitors.

Acknowledgments

The authors acknowledge the support from the Natural Science Foundation, in the Fujian Province,

China (Grant No. 2016J01729), and the support from the Key Program of the Youth Natural

Science Foundation of the Fujian Province University, China (Grant No. JZ160413).

Appendix A. Supplementary data

Supplementary data associated with this study can be found in the Supporting Materials.

References

[1] L. Lou, B. Wu, L. Wang, L. Luo, X. Xu, J. Hou, B. Xun, B. Hu, Y. Chen, Bioresour. Technol.

21 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

2011, 102, 4036–4061.

[2] M. Uchimiya, L.H. Wartelle, K.T. Klasson, C.A. Fortier, I.M. Lima, J. Agric. Food. Chem. 2011,

59, 2501–2510.

[3] X. Wu, L. Jiang, C. Long, Z. Fan, Nano Energy 2015, 13, 527–536.

[4] M. Biswal, A. Banerjee, M. Deo, S. Ogale, Energy Environ. Sci. 2013, 6, 1249–1259.

[5] J.-R. Wang, F. Wan, Q.-F. Lü, F. Chen, Q. Lin, J. Mater. Sci.Technol. 2018, 34, 1959–1968.

Accepted Manuscript
[6] Q. Liang, L. Ye, Z.-H. Huang, Q. Xu, Y. Bai, F. Kang, Q.-H. Yang, Nanoscale 2014, 6, 13831–

13837.

[7] a) C. Liu, F. Li, L.P. Ma, H.M. Cheng, Adv. Mater. 2010, 22, 28–62; b) J.-R. Wang, F. Wan, Q.-

F. Lü, F. Chen, Q. Lin, J. Mater. Sci. Technol. 2018, 34,1959–1968.

[8] M. Inagaki, H. Konno, O. Tanaike, J. Power Sources 2010, 195, 7880–7903.

[9] W.-H. Li, K. Ding, H.-R. Tian, M.-S. Yao, B. Nath, W.-H. Deng, Y. Wang, G. Xu, Adv. Funct.

Mater. 2017, 27, 1702067.

[10] F. Wang, X. Wu, X. Yuan, Z. Liu, Y. Zhang, L. Fu, Y. Zhu, Q. Zhou, Y. Wu, W. Huang, Chem.

Soc. Rev. 2017, 46, 6816–6854.

[11] Y. Liao, H. Wang, M. Zhu, A. Thomas, Adv. Mater. 2018, 30, 1705710.

[12] H. Wang, Z. Cheng, Y. Liao, J. Li, J. Weber, A. Thomas, C.F.J. Faul, Chem. Mater. 2017, 29,

4885–4893.

[13] S. Xiong, F. Yang, H. Jiang, J. Ma, X. Lu, Electrochim. Acta 2012, 85, 235–242.

[14] T.-T. Lin, W.-H. Lai, Q.-F. Lü, Y. Yu, Electrochim. Acta 2015, 178, 517–524.

[15] X.-Q. Lin, W.-D. Wang, Q.-F. Lü, Y.-Q. Jin, Q. Lin, R.Liu, J. Mater. Sci.Technol. 2017, 33,

1339–1345.

[16] L.T. Le, M.H. Ervin, H. Qiu, B.E. Fuchs, W.Y. Lee, Electrochem. Commun. 2011,13, 355–358.
22 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

[17] Z.-J. Qiao, M.-M. Chen, C.-Y. Wang, Y.-C. Yuan, Bioresour. Technol. 2014, 163, 386–389.

[18] L. Hao, X. Li, L. Zhi, Adv. Mater. 2013, 25, 3899–3904.

[19] T.-C. Chou, C.-H. Huang, R.-A. Doong, Synth. Met. 2014, 194, 29–37.

[20] P. Cheng, S. Gao, P. Zang, X. Yang, Y. Bai, H. Xu, Z. Liu, Z. Lei, Carbon 2015, 93, 315–324.

[21] Y. Li, G. Wang, T. Wei, Z. Fan, P. Yan, Nano Energy 2016,19, 165–175.

[22] Y. Qu, G. Zan, J. Wang, Q. Wu, J. Mater. Chem. A 2016, 4, 4296–4304.

Accepted Manuscript
[23] Y. Lv, L. Gan, M. Liu, W. Xiong, Z. Xu, D. Zhu, D.S. Wright, J. Power Sources 2012, 209,

152–157.

[24] W.H. Qu, Y.Y. Xu, A.H. Lu, X.Q. Zhang, W.C. Li, Bioresour. Technol. 2015, 189, 285–291.

[25] R.-J. Mo, Y. Zhao, M. Wu, H.-M. Xiao, S. Kuga, Y. Huang, J.-P. Li, S.-Y. Fu, RSC Adv. 2016,

6, 59333–59342.

[26] J. Zhou, Z. Sheng, H. Han, M. Zou, C. Li, Mater. Lett. 2012, 66, 222–224.

[27] R.J. White, V. Budarin, R. Luque, J.H. Clark, D.J. Macquarrie, Chem. Soc. Rev. 2009, 38,

3401–3418.

[28] R.J. White, C. Antonio, V.L. Budarin, E. Bergström, J. Thomas-Oates, J.H. Clark, Adv. Funct.

Mater. 2010, 20, 1834–1841.

[29] S.T. Senthilkumar, R.K. Selvan, N. Ponpandian, J.S. Melo, Y.S. Lee, J. Mater. Chem. A 2013,

1, 7913–7919.

[30] K.V. Sankar, R. Kalai Selvan, Carbon 2015, 90, 260–273.

[31] S. Roldán, M. Granda, R. Menéndez, R. Santamaría, C. Blanco, Electrochim. Acta 2012, 83,

241–246.

[32] H. Yu, J. Wu, L. Fan, S. Hao, J. Lin, M. Huang, J. Power Sources 2014, 248, 1123–1126.

[33] Z.J. Zhang, X.Y. Chen, J. Electroanal. Chem. 2016, 764, 45–55.
23 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

[34] X.Q. Li, P.F. Zhang, W.D. Duan, D.L. Zhang, C. Li, J. Chin. Medic. Mater. 2009, 32, 1227–

1229.

[35] A.E. Ismanto, S. Wang, F.E. Soetaredjo, S. Ismadji, Bioresour. Technol. 2010, 101, 3534–3540.

[36] Y. Lv, F. Zhang, Y. Dou, Y. Zhai, J. Wang, H. Liu, Y. Xia, B. Tu, D. Zhao, J. Mater. Chem.

2012, 22, 93–99.

[37] R.K. Sharma, W.G. Chan, M.R. Hajaligol, J. Anal. Appl. Pyrolysis 2009, 86, 122–134.

Accepted Manuscript
[38] E.B. Sanders, A.I. Goldsmith, J.I. Seeman, J. Anal. Appl. Pyrolysis 2003, 66, 29–50.

[39] Y. Zhao, J. Qiu, H. Feng, M. Zhang, L. Lei, X. Wu, Chem. Eng. J. 2011, 173, 659–666.

[40] A. Jain, R. Balasubramanian, M.P. Srinivasan, Chem. Eng. J. 2016, 283, 789–805.

[41] C.-W. Tsai, M.-H. Tu, C.-J. Chen, T.-F. Hung, R.-S. Liu, W.-R. Liu, M.-Y. Lo, Y.-M. Peng, L.

Zhang, J. Zhang, D.-S. Shy, X.-K. Xing, RSC Adv. 2011, 1, 1349–1357.

[42] Y. Cai, Y. Luo, H. Dong, X. Zhao, Y. Xiao, Y. Liang, H. Hu, Y. Liu, M. Zheng, J. Power

Sources 2017, 353, 260–269.

[43] K. Ding, Q. Liu, Y. Bu, Y. Huang, J. Lv, J. Wu, S.C. Abbas, Y. Wang, RSC Adv. 2016, 6,

93318–93324.

[44] M. Sevilla, A.B. Fuertes, Carbon 2013, 56, 155–166.

[45] K. Wang, N. Zhao, S. Lei, R. Yan, X. Tian, J. Wang, Y. Song, D. Xu, Q. Guo, L. Liu,

Electrochim. Acta 2015, 166, 1–11.

[46] Y. Zhu, S. Murali, M.D. Stoller, K.J. Ganesh, W. Cai, P.J. Ferreira, A. Pirkle, R.M. Wallace,

K.A. Cychosz, M. Thommes, Science 2011, 332, 1537–1541.

[47] J. Zhang, J. Zhou, D. Wang, L. Hou, F. Gao, Electrochim. Acta 2016, 191, 933–939.

[48] G. Ma, Q. Yang, K. Sun, H. Peng, F. Ran, X. Zhao, Z. Lei, Bioresour. Technol. 2015, 197,

137–142.
24 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

[49] A.G. Pandolfo, A.F. Hollenkamp, J. Power Sources 2006, 157, 11–27.

[50] D. Qu, J. Power Sources 2002, 109, 403–411.

[51] a) D. Puthusseri, V. Aravindan, S. Madhavi, S. Ogale, Energy Environ. Sci. 2014, 7, 728–735;

b) G. Wang, J. Zhang, S. Kuang, J. Zhou, W. Xing, S. Zhuo, Electrochim. Acta 2015, 153,

273–279.

[52] a) B. Xu, S. Yue, Z. Sui, X. Zhang, S. Hou, G. Cao, Y. Yang, Energy Environ. Sci. 2011, 4,

Accepted Manuscript
2826–2830; b) Y. Zhou, J. Ren, L. Xia, Q. Zheng, J. Liao, E. Long, F. Xie, C. Xu, D. Lin,

Electrochim. Acta 2018, 284, 336–345.

[53] Y. Huang, P. Wu, Y. Wang, W. Wang, D. Yuan, J. Yao, J. Mater. Chem. A 2014, 2, 19765–

19770.

[54] Q. Liu, Y. Wang, L. Dai, J. Yao, Adv. Mater. 2016, 28, 3000–3006.

[55] Y. Liu, Z. Shi, Y. Gao, W. An, Z. Cao, J. Liu, ACS Appl. Mater. Interfaces 2016, 8, 28283–

28290.

[56] H. Wang, Z. Xu, A. Kohandehghan, Z. Li, K. Cui, X. Tan, T.J. Stephenson, C.K. King'Ondu,

C.M. Holt, B.C. Olsen, ACS Nano 2013, 7, 5131–5141.

[57] Y.S. Yun, C. Im, H.H. Park, I. Hwang, Y. Tak, H.-J. Jin, J. Power Sources 2013, 234, 285–291.

[58] J. Chang, Z. Gao, X. Wang, D. Wu, F. Xu, X. Wang, Y. Guo, K. Jiang, Electrochim. Acta 2015,

157, 290–298.

[59] A. Elmouwahidi, Z. Zapata-Benabithe, F. Carrasco-Marin, C. Moreno-Castilla, Bioresour.

Technol. 2012,111, 185–190.

[60] X. Du, W. Zhao, Y. Wang, C. Wang, M. Chen, T. Qi, C. Hua, M. Ma, Bioresour. Technol.

2013, 149, 31–37.

[61] Q. Cheng, J. Tang, J. Ma, H. Zhang, N. Shinya, L.-C. Qin, Carbon 2011, 49, 2917–2925.
25 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

[62] H.-M. Luo, Y.-F. Yang, Y.-X. Sun, X. Zhao, J.-Q. Zhang, J. Solid State Electrochem. 2015, 19,

1491–1500.

[63] R. Farma, M. Deraman, A. Awitdrus, I.A. Talib, E. Taer, N.H. Basri, J.G. Manjunatha, M.M.

Ishak, B.N.M. Dollah, S.A. Hashmi, Bioresour. Technol. 2013, 132, 254–261.

[64] S.-H. Du, L.-Q. Wang, X.-T. Fu, M.-M. Chen, C.-Y. Wang, Bioresour. Technol. 2013, 139,

406–409.

Accepted Manuscript
[65] H.L. Ji, N. Park, B.G. Kim, D.S. Jung, K. Im, J. Hur, J.W. Choi, ACS Nano 2013, 7, 9366–

9374.

[66] J. Wu, H. Yu, L. Fan, G. Luo, J. Lin, M. Huang, J. Mater. Chem. 2012, 22, 19025–19030.

[67] T. Brousse, D. Bélanger, J. W. Long, J. Electrochem. Soc. 2015, 162, A5185–A5189.

[68] H. Yu, L. Fan, J. Wu, Y. Lin, M. Huang, J. Lin, Z. Lan, RSC Adv. 2012, 2, 6736–6740.

26 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

Table List:

Table 1. Elemental analyses of the biocarbon samples.

Table 2. Contents of different forms of nitrogen and oxygen in the biocarbon samples obtained from

XPS analysis.

Table 3. Characteristics of the as-obtained biocarbon porous structures.

Accepted Manuscript
Figure captions:

Scheme 1. Schematic illustration of ACW-2 preparation as an electrode.

Figure 1. FE-SEM images of (a) CW-700, (b) ACW-1, (c) ACW-2, and (d) ACW-3.

Figure 2. TGA-DTG curves (a) of watermelon rind, and FT-IR spectra (b), XRD diffraction curves

(c), and Raman spectra (d) of biocarbon samples.

Figure 3. (a) XPS survey spectrum, (b) C 1s, (c) N 1s, (d) O 1s of ACW-2; and (e) N 1s, (f) O 1s

high-resolution spectra of ACW-1, ACW-2 and ACW-3.

Figure 4. (a) N2 adsorption and desorption isotherms, (b) micropore size distributions obtained by

using HK method, and (c) mesopore size distributions obtained by using BJH method of obtained

samples.

Figure 5. (a) CV curves of biocarbons at scan rate of 50 mV s-1, and (b) GCD curves of biocarbons

at current density of 1 A g-1.

Figure 6. (a) CV curves of ACW-2 at different scan rates, (b) GCD curves of ACW-2, and (c) specific

capacitances of four biocarbons at different current densities.

Figure 7. (a) Nyquist plots and (b) cycling stabilities of biocarbons at 100 mV s-1; and (c,d) FE-

SEM images of ACW-2 electrode after 5000 cycling test at 100 mV s-1.
27 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

Figure 8. (a) CV curves of ACW-2 at scan rate of 50 mV s-1, and (b) GCD curves of ACW-2 at

current density of 1 A g-1 in 6 M KOH of electrolytes with different concentration of PPD.

Figure 9. GCD curves of ACW-2//ACW-2 devices at different current densities in (a) 6 M KOH and

(b) 6 M KOH + 1 mM PPD; (c) ragone plots of symmetrical supercapacitor.

Accepted Manuscript

28 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

Accepted Manuscript
Scheme 1. Schematic illustration of ACW-2 preparation as an electrode.

(a)  (b) 

(c)  (d) 

Figure 1. FE-SEM images of (a) CW-700, (b) ACW-1, (c) ACW-2, and (d) ACW-3.

29 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

100
(a) (b)

Transmittance (a.u.)
18.09 %
80 CW-700
Weight (%)

DTG
60 34.52 % ACW-1

ACW-2
40 10.37 %

Accepted Manuscript
TGA ACW-3
20

0
0 200 400 600 800 10001200 4000 3200 2400 1600-1
800
Temperature (C) Wavenumber (cm )

(002) (c) D G (d)


CW-700
Intensity (a.u.)
Intensity (a.u.)

(100) CW-700

ACW-1 ACW-1

ACW-2 ACW-2

ACW-3
ACW-3

20 40 60 80 1000 2000 -1 3000


2degree Raman Shift (cm )

Figure 2. TGA-DTG curves (a) of watermelon rind, FT-IR spectra (b), XRD diffraction curves (c),

and Raman spectra (d) of biocarbon samples.

30 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

(a) C 1s (b)
C-C
Intensity (a.u.)

Intensity (a.u.)
O 1s C-O
C=O
HO-C=O
N 1s C-N

Accepted Manuscript
C=C

800 600 400 200 0 294 291 288 285 282 279
Binding Energy (eV) Binding Energy (eV)

Graphitic-N (d)
(c) C-O
Pyrrolic-N
Intensity (a.u.)

Intensity (a.u.)

Pyridinic-N C=O
-OH

404 402 400 398 396 538 536 534 532 530 528 526
Binding Energy (eV) Binding Energy (eV)

(e) Graphitic-N Pyrrolic-N C-O


(f) -OH
ACW-1 Pyridinic-N C=O
ACW-1
Intensity (a.u.)
Intensity (a.u.)

ACW-2
ACW-2

ACW-3 ACW-3

406 404 402 400 398 396 536 534 532 530 528
Binding Energy (eV)
Binding Energy (eV)  

Figure 3. (a) XPS survey spectrum, (b) C 1s, (c) N 1s, (d) O 1s of ACW-2; and (e) N 1s, (f) O 1s

high-resolution spectra of ACW-1, ACW-2 and ACW-3.

31 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

dV/dw Pore Volume (cm /g nm)


Quantity Adsorbed (cm /g STP)

500 (a) 1.2 (b)


400 ACW-1

3
3

ACW-2
300 0.8
ACW-3
ACW-1 CW-700
200
0.4

Accepted Manuscript
ACW-2
100 ACW-3
0 CW-700 0.0

0.0 0.2 0.4 0.6 0.8 1.0 0 1 2 3


Relative Pressure (P/P0) Pore Width (nm)
dV/dw Pore Volume (cm /g nm)

0.3 (c)
ACW-1
3

ACW-2
0.2
ACW-3
CW-700
0.1

0.0

1 10 100
Pore Width (nm)

Figure 4. (a) N2 adsorption and desorption isotherms, (b) micropore size distributions obtained by

using HK method, and (c) mesopore size distributions obtained by using BJH method of obtained

samples.

32 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

20 CW-700 ACW-2 0.0 CW-700


(a) (b)
Current Density (A g )
-1
15 ACW-1 ACW-3 ACW-1

Potential (V)
10 -0.2 ACW-2
5 -0.4 ACW-3

0 -0.6
-5
-0.8
-10
-15 -1.0
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0 100 200 300 400 500
Time (s)

Accepted Manuscript
Potential (V)

Figure 5. (a) CV curves of biocarbons at scan rate of 50 mV s-1, and (b) GCD curves of biocarbons

at current density of 1 A g-1.

100
Current Density (A g )

(a) 50 mVs-1 0.0 (b) -1


-1

5 mVs-1 1Ag
80 100 mVs-1 -1
20 mVs-1 2Ag
Potential (V)

60 300 mVs-1 -0.2 5Ag


-1

40 10 A g
-1
-0.4
20
0 -0.6
-20 -0.8
-40
-60 -1.0

-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0 100 200 300 400 500 600
Potential (V) Time (s)
300
(c) ACW-1 ACW-3
Capacitance (F g )

ACW-2 CW-700
-1

200

100

0
0 2 4 6 8 10
-1
Current Density (A g )

Figure 6. (a) CV curves of ACW-2 at different scan rates, (b) GCD curves of ACW-2, and (c)

specific capacitances of four biocarbons at different current densities.

33 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

60 ACW-1
(a) 120 (b)

Retention Rate(%)
ACW-2
100
ACW-3
80
-Z'' ()

CW-700 ACW-1
6
30 5
60 ACW-2
4
ACW-3
3 40
2 CW-700
1 20
0
0 1 2 3 4 5 6
0 0
0 30 60 0 1250 2500 3750 5000

Accepted Manuscript
Z' () Cycle Number

(c)  (d) 

Figure 7. (a) Nyquist plots and (b) cycling stabilities of biocarbons at 100 mV s-1; and (c,d) FE-

SEM images of ACW-2 electrode after 5000 cycling test at 100 mV s-1.
40
Current Density (A g )

(a) (b)
-1

0.0 KOH
KOH+0.2mM PPD
Potential (V)

20 -0.2 KOH+1.0mM PPD


KOH+1.5mM PPD
-0.4
0
-0.6
KOH
-20 KOH+0.2mM PPD -0.8
KOH+1.0mM PPD
KOH+1.5mM PPD -1.0
-40
-1.0 -0.8 -0.6 -0.4 -0.2 0.0 0 300 600 900 1200 1500
Potential (V) Time (s)

Figure 8. (a) CV curves of ACW-2 at scan rate of 50 mV s-1, and (b) GCD curves of ACW-2 at

current density of 1 A g-1 in 6 M KOH of electrolyte with different PPD concentration.

34 
 

This article is protected by copyright. All rights reserved.


Energy Technology 10.1002/ente.201800628

0.0 (a)

Accepted Manuscript
6M KOH
-1 0.0 (b) 6M KOH +1mM PPD
0.5A g -1
0.5A g
Potential (V)

-0.2 -1
-0.2

Potential (V)
1A g -1
-1 1A g
2A g -1
-0.4 -1 -0.4 2A g
5A g -1
5A g
-0.6 -0.6
-0.8 -0.8
-1.0 -1.0
0 50 100 150 200 0 50 100 150 200 250 300
Time (s) Time (s)
100
Energy density (Wh kg )
-1

(c)

10

6M KOH
1 6M KOH+1mM PPD

0.1
100 1000 -1 10000
Power density (W kg )

Figure 9. GCD curves of ACW-2//ACW-2 devices at different current densities in (a) 6 M KOH and

(b) 6 M KOH + 1 mM PPD; (c) ragone plots of symmetrical supercapacitor.

35 
 

This article is protected by copyright. All rights reserved.

You might also like