Download as pdf or txt
Download as pdf or txt
You are on page 1of 49

Materials Research Bulletin

Efficient Production of Activated Carbons from PET for Organic Supercapacitor


Applications: A Single-Step Approach
--Manuscript Draft--

Manuscript Number: MRB-D-24-01144

Article Type: Research Paper

Section/Category: Electrochemistry(incl. electrocatalysis, photocatalysis, batteries, electrochemical


sensors & DSSCs)

Keywords: Single-step method; activated carbon; polyethylene terephthalate; supercapacitor;


Organic electrolyte

Abstract: This study introduces a novel approach for synthesizing supercapacitor electrode
materials using polyethylene terephthalate (PET), a plastic facing recycling challenges.
Unlike conventional methods that involve separate carbonization and activation steps,
we employ a single-step method. In this process, activated carbons are synthesized
using potassium hydroxide (KOH) as an activating agent. Remarkably, the yield via the
single-step method (S_PACXs) exceeds that of the two-step method (T_PACYs) by at
least 1.7 times. During the single-step process, KOH forms a layer on the PET surface
before activation, which leads to increased yields, higher specific surface areas, and
more developed mesopores. Furthermore, S_PACXs exhibit superior specific surface
areas compared to commercial activated carbon. These enhanced properties
significantly improve electrochemical performance, with S_PACXs demonstrating
superior performance compared to T_PACYs. Ultimately, this study validates the
efficiency of the single-step method in producing high-quality activated carbon from
PET, saving time and energy, and outperforming the two-step method.

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Highlights

Highlights

► Polyethylene terephthalate (PET) chips are direct converted into activated carbons.

► The single-step method, which simultaneously carries out carbonization and activation, is

explained.

► PET-derived activated carbon synthesized in the single step exhibits superior physical

properties.

► The single-step method is deemed suitable for synthesizing electrode materials for high-

performance supercapacitors.

► The method reported in this study represents possible a solution for the recycling of PET.
Graphical Abstract
Manuscript File Click here to view linked References

Materials Research Bulletin

Efficient Production of Activated Carbons from PET for Organic


Supercapacitor Applications: A Single-Step Approach

Jongyun Choia, Wonjong Jungb*, Ji Chul Junga**

a
Department of Chemical Engineering, Myongji University,
Yongin 17058, Republic of Korea

b
Department of Mechanical, Smart, and Industrial Engineering, Gachon University,
Seongnam 13120, Republic of Korea

*Corresponding author
(E-mail: wonjongjung@gachon.ac.kr)

**Corresponding author
(E-mail: jcjung@mju.ac.kr)
Abstract
This study introduces a novel approach for synthesizing supercapacitor electrode

materials using polyethylene terephthalate (PET), a plastic facing recycling challenges. Unlike

conventional methods that involve separate carbonization and activation steps, we employ a

single-step method. In this process, activated carbons are synthesized using potassium

hydroxide (KOH) as an activating agent. Remarkably, the yield via the single-step method

(S_PACXs) exceeds that of the two-step method (T_PACYs) by at least 1.7 times. During the

single-step process, KOH forms a layer on the PET surface before activation, which leads to

increased yields, higher specific surface areas, and more developed mesopores. Furthermore,

S_PACXs exhibit superior specific surface areas compared to commercial activated carbon.

These enhanced properties significantly improve electrochemical performance, with S_PACXs

demonstrating superior performance compared to T_PACYs. Ultimately, this study validates

the efficiency of the single-step method in producing high-quality activated carbon from PET,

saving time and energy, and outperforming the two-step method.

Keywords: Single-step method; activated carbon; polyethylene terephthalate; supercapacitor;

Organic electrolyte
1. Introduction

In recent years, energy storage has emerged as a critical concern, driven by the expanding

eco-friendly electric vehicle market and the increasing interest in energy storage systems (ESS)

for efficient energy storage and management [1–5]. Various energy storage devices, including

secondary batteries, fuel cells, and supercapacitors, find applications across diverse fields

owing to their unique advantages. Unlike conventional Li-ion batteries that store energy

through redox reactions, supercapacitors rely on the physical adsorption of electrolyte ions for

energy storage [5–9]. Renowned for their high-power density, rapid charge–discharge

capability, and prolonged lifespan, supercapacitors serve as ideal complements to secondary

batteries, especially in applications where a high-power density is paramount, such as mobile

devices and electric vehicles [8–12].

Despite their advantages, supercapacitors exhibit low energy density characteristics,

prompting ongoing research efforts aimed at enhancing their energy density [13–16]. The key

to achieving this lies in maximizing the contact area between the electrode and electrolyte ions,

which is crucial for efficient energy storage [16–19]. One promising approach involves coating

the electrode current collector with an active material known for its porous structure and high

specific surface area. Activated carbon stands out as a prime candidate due to its ability to

provide ample space for electrolyte ion adsorption, thereby boosting energy density [15, 19–

21]. Additionally, optimizing the pore size distribution in activated carbons enhances

electrolyte ion mobility, while improving electrical conductivity, influenced by the crystallinity

of carbon materials, further enhances supercapacitor performance [21–23].

Various carbon-containing materials, including wood, coconut shells, coke, fallen leaves,

plastics, etc., serve as raw materials for producing activated carbon. Among these, plastics,

notably polyethylene terephthalate (PET), have garnered interest due to their widespread use
and significant environmental impact [24–30]. PET, commonly found in water and beverage

bottles, contributes substantially to plastic waste and environmental pollution. Disposal

methods such as incineration and landfill exacerbate these issues, underscoring the urgency to

explore recycling options and create higher value-added products from plastic waste. With

carbon comprising approximately 62.5 wt% of a PET molecule, recycling PET into activated

carbon presents a promising avenue for addressing plastic waste concerns.

Traditionally, manufacturing activated carbon from raw materials involves a two-step heat

treatment process encompassing carbonization and activation, typically conducted at

temperatures exceeding 600°C [31–33]. Previous research has highlighted the challenges

associated with this approach, including high energy consumption, prolonged processing times,

and low yields [34,35]. To address these limitations, this study proposes a novel single-step

heat treatment method for producing PET-derived activated carbon for supercapacitors. This

method offers enhanced efficiency by saving energy, increasing yields, and reducing

production times compared to the conventional two-step method.

In this study, we synthesized PET-derived activated carbons at various temperatures using

the single-step method, which concurrently conducts carbonization and activation, as depicted

in Fig. 1. For comparison, PET-derived activated carbons were also synthesized using the

conventional two-step method. In both methods, potassium hydroxide (KOH) was employed

as the activating agent. The comparative analysis of yield and physical properties of the

synthesized carbon materials, utilizing techniques such as Raman spectroscopy and N2

adsorption–desorption isotherms, provides insights into the efficacy of the synthesis methods.

Based on the physical properties, we assess the suitability of the synthesized carbon materials

as electrode materials for supercapacitors. Finally, we evaluate the electrochemical

characteristics by incorporating the synthesized carbon materials into coin-type organic


electrolyte supercapacitors.

Fig.1. Scheme of PET-derived activated carbons fabrication process.

2. Experimental

2.1. Synthesis of PET-derived Carbon Materials

To synthesize PET-derived activated carbons through the single-step method, wherein

carbonization and activation processes occur simultaneously, PET chips (Lotte Chemical,

Republic of Korea) were initially pulverized into small sizes using a blender (CB15VE, Waring

Commercial, USA). The pulverized PET was then mixed with potassium hydroxide (KOH) at

a mass ratio of 1:4. The resulting mixture was placed in an alumina boat and subjected to heat

treatment in a tubular furnace under a nitrogen atmosphere to produce activated carbons. The

heating process maintained a fixed rate of 5°C min−1, with the target temperature held for 4

hours. Target temperatures ranged from 600 to 900°C. After heat treatment, the product was

naturally cooled to room temperature, followed by thorough washing with distilled water to

neutralize the remaining KOH. Subsequently, the washed product was dried in an oven at
100°C for 24 hours to eliminate moisture, yielding the final product. Activated carbon

synthesized via the single-step method was denoted as S_PACX, where X represents the heat

treatment temperature.

For comparative purposes, PET-derived activated carbons were synthesized using the

conventional two-step method, wherein carbonization and activation occur sequentially.

Initially, the pulverized PET was carbonized at 800°C for 2 hours in a tubular furnace, yielding

PET-derived carbons designated as PC800. Subsequently, the activation process involved

mixing PC800 with KOH in a mass ratio of 1:4 and subjecting the mixture to heat treatment in

a tubular furnace. Target activation temperatures ranged from 600 to 900°C, with a duration of

2 hours. The heating conditions, gas atmosphere, and flow rate remained consistent with those

of the single-step method. Washing and drying procedures were identical to those of the single-

step method. PET-derived activated carbon synthesized via the two-step method was

designated as T_PACY, where Y indicates the activation temperature. Additionally,

commercially available activated carbon MSP20 for supercapacitors was purchased and

utilized without further treatment for comparison purposes.

2.2. Characterization of Carbon Materials

Physical properties of the activated carbons (S_PACXs and T_PACYs) including specific

surface area, average pore diameter, and pore volume were determined through N2 adsorption–

desorption isotherm analysis using a constant-volume adsorption apparatus (BELSORP-max

II, MicrotracBEL, Japan) at 77K. The specific surface area was calculated using the Brunauer–

Emmett–Teller (BET) equation [36]. The crystallinity of the carbon materials was assessed via

Raman spectroscopy (DXP2Xi, Thermo, USA) within the range of 50 cm−1 to 3400 cm−1.

Scanning electron microscopy (SEM) (SU-70, Hitachi, Japan) was employed to examine the
morphology of PET-derived carbon materials (PET800, PACXs, S_PACXs). The thermal

properties of PET were analyzed using thermal gravimetric analysis (TGA) conducted with a

high-temperature simultaneous thermal analyzer (STA449 F3, NETZSCH, Germany).

2.3. Fabrication of Supercapacitors

Prepared activated carbons (S_PACXs, T_PACYs, and MSP20) served as the active

material, while Super-P (M.M.M. Carbon Co., Belgium) acted as the conductive material, and

polyvinylidene fluoride (PVDF) dissolved in 1-methyl-2-pyrrolidone (NMP, Daejung Chem.,

Republic of Korea) functioned as the binder. The active material, conductive material, and

binder were mixed in a ratio of 8:1:1. The mixture underwent ball milling for 30 minutes,

followed by the addition of a binder. To regulate the viscosity of the mixture, a small quantity

of NMP was added, and stirring continued for over 1 hour. The resulting slurry was then coated

onto an aluminum current collector using a doctor blade to achieve a thickness of 250 μm.

Following coating, the electrode underwent drying in an oven at 70°C for 24 hours and

subsequently in a vacuum oven at 70°C for another 24 hours. The dried electrode was

compressed via roll pressing at 80°C and punched into circular electrodes with a diameter of

16 mm. Subsequently, coin-type supercapacitors (CR2032) were assembled within a glove box

filled with Argon gas. The electrolyte, 1-M tetraethylammonium tetrafluoroborate in

acetonitrile (TABF4/ACN, Enchem, Republic of Korea), was applied to the electrodes and

allowed to be immersed for 12 hours before assembly. Finally, the assembled supercapacitors

were completely sealed using a crimper (Wellcos Co., Republic of Korea). Fig. 2 shows the

assembly process of a supercapacitor and the synthesis of activated carbon.


2.4. Electrochemical Characterization of Fabricated Supercapacitors

Electrochemical properties of the fabricated supercapacitors were analyzed through cyclic

voltammetry (CV) and galvanostatic charge–discharge (C–D) analysis conducted using a

multi-channel battery charge–discharge tester (WBCS3000, WonATech, Republic of Korea).

CV curves were recorded within a voltage range of 0–2.7 V at scan rates ranging from 5 to 300

mV s−1, while C–D analysis was conducted across a current density range of 0.1–7 A g−1 within

the same voltage range. Furthermore, to assess long-term durability, C–D analysis was repeated

3000 times at a current density of 1 A g−1. Resistance characteristics of the supercapacitor were

determined via electrochemical impedance spectroscopy (EIS) analysis using an

electrochemical workstation (CS310, CorrTset, China) over a frequency range from 100 kHz

to 0.01 kHz, presented as a Nyquist plot. Electrode material resistance analysis was also

performed by coating the electrode material onto a non-conducting OHP film and measuring it

using a four-point probe (FPP, CMT-SR1000N, Advanced Instrument Technology, Republic of

Korea).

Fig. 2. Scheme of the fabrication process for supercapacitor.


3. Result and discussion
3.1. Single-step heat treatment mechanism
The synthesis of activated carbon necessarily involves heat treatment processes under an

inert atmosphere, typically conducted at temperatures of around 600°C or higher [37,38]. To

evaluate the thermal properties of PET used in activated carbon production, we conducted

thermogravimetric analysis (TGA). Fig. 3 illustrates the TGA analysis results of the PET

utilized in this experiment, carried out under an inert atmosphere using nitrogen. Initially, the

mass of PET exhibited minimal changes until it reached approximately 370°C. Subsequently,

a rapid mass decrease occurred between 370°C and 470°C, attributed to the thermal

decomposition of polymers. Upon reaching 800°C, only about 15% of the initial mass remained

as carbon material. This low yield poses a challenge for the production of PET-derived carbon

materials.

Fig. 3. TGA date of PET obtained under nitrogen condition.

In this study, we introduced a single-step heat treatment method aimed at enhancing the

yield of PET-derived activated carbon while simplifying the production process. To


demonstrate the efficiency of the single-step method in increasing product yield compared to

the traditional two-step approach, we conducted experiments using both methods and

compared the final product yields. The yield of PET-derived activated carbon was calculated

using Eqs. (1) and (2):

𝑦𝑡𝑤𝑜−𝑠𝑡𝑒𝑝 = 𝑚𝑡_𝑎𝑐 /𝑚𝑃𝐸𝑇 × 100% (1)

𝑦𝑠𝑖𝑛𝑔𝑙𝑒−𝑠𝑡𝑒𝑝 = 𝑚𝑠_𝑎𝑐 /𝑚𝑃𝐸𝑇 × 100% (2)

Here, ytwo-setp and ysingle-step represent the yields of activated carbons derived from PET through

the two-step and single-step methods, respectively. mt_ac denotes the weight of activation

carbon obtained by the two-step method, and ms_ac represents the weight of activated carbon

obtained by the single-step method. mPET denotes the weight of PET used in the two-step and

single-step methods. The resulting yields are illustrated in Fig. 4.

The single-step heat treatment method significantly increased the yield of PET-derived

activated carbons, achieving a yield at least 1.7 times higher under the same activation

temperature conditions compared to the two-step method. For instance, the yield of S_PAC600

was 25.5%, higher than the yield of T_PAC600 (13.9%) heat-treated at the same activation

temperature. This represented the highest yield among all samples. Notably, SPAC800,

produced in a single step at 800°C, exhibited a higher yield (15.3%) compared to T_PAC600

(13.9%) heat-treated at an even lower temperature, which had the highest yield among activated

carbons produced using the two-step method. Furthermore, the single-step method enables

simultaneous carbonization and activation, leading to more efficient use of energy and time

compared to the two-step method. Thus, the single-step method not only efficiently utilizes

time and energy by simplifying the fabrication process but also yields a higher quantity of final

products.
Fig. 4. Yield of PET-derived activated carbons.

Interestingly, we observed a unique result where synthesizing activated carbon at 600, 700,

and 800 °C using the single-step method yielded more carbon materials compared to the TGA

results. This suggests that the improved yield is not solely due to the simplification of the

synthesis method but also includes additional influences. We identified KOH, added as an

active agent, as the main factor for the yield increase. Consequently, we aim to clarify the cause

of this unique result by investigating the single-step reaction mechanism involving KOH and

PET.

The reaction mechanism of the single-step method was divided and analyzed into three

stages for analysis, based on temperature changes, as illustrated in Fig. 5. The first stage

involves heating from room temperature to 400°C The melting point of KOH is 361°C, and as

the temperature rises to 400°C, the excess KOH melts and forms a layer that fully covers the

PET surface. This layer facilitates the thermal decomposition of PET, typically around 370°C.
The second stage, between 400°C and 600°C, involves intensive PET decomposition,

transitioning into carbon materials. Numerous cracks appear inside the PET particles, allowing

KOH to infiltrate these cracks. This stage also involves the activation reaction by KOH in the

partially converted carbon materials. The final stage is the KOH activation reaction at high

temperatures above 600°C, occurring simultaneously on the surface and inside due to the

presence of KOH on the surface and infiltrated KOH within.

In the single-step process, the formation of a KOH layer on the PET surface is key. This

occurs when KOH, melting at lower temperatures, completely covers the PET. This KOH layer

provides a conducive environment for the transition of PET into carbon materials. Generally,

during the carbonization process of PET, lighter gaseous components like CO and CO2 are

formed, along with relatively heavier polyaromatic hydrocarbons (PAHs) transitioning into

carbon [27,39]. In the single-step process, the mobility of gases generated by PET

decomposition is limited due to the KOH layer formed on the surface. Large molecules like

PAHs cannot pass through the KOH layer and remain, contributing to favorable conditions for

synthesizing carbon materials and increasing the quantity of the final product. In other words,

through the formation of a KOH layer on the PET surface, the single-step method can overcome

the challenge of low yield in PET-derived carbon materials.

Fig. 5. Scheme of the single-step reaction process.


3.2. Morphology

The morphological transformation of PET into activated carbons via heat treatment was

examined in this section. We observed that the resulting PET-derived activated carbons

exhibited morphology predominantly influenced by the synthesis method rather than the heat

treatment temperature. Our analysis focused on comparing S_PAC700 and T_PAC700, which

suitably illustrate the variances depending upon the synthesis method. SEM images depicting

PET and the synthesized activated carbons are presented in Fig. 6. Notably, the particle size of

S_PAC700 measures in the tens of micrometers. This is markedly different from that of

T_PAC700, which has a particle size of less than 10 micrometers, as distinguished in Fig. 6(b)

and Fig. 6(c). The smaller particle sizes of T_PAC700 could be attributed to the occurrence of

cracking and subsequent fragmentation of particles induced by gas evolution during the

synthesis process. In contrast, S_PAC700 retains relatively larger particle sizes owing to the

encapsulation of PET particles by the KOH layer formed on the PET surface during the single-

step process. This encapsulation mechanism prevents fragmentation even when cracks occur.

Therefore, it is evident that the KOH layer formed during the single-step process significantly

influences the particle size of the activated carbons. Moreover, we presume that these changes

in particle size could potentially induce alterations in not only the yield of the activated carbons

but also other properties of the activated carbons.

Fig. 6. SEM images of (a) PET, (b) S_PAC700, and (c) T_PAC700.
3.3 Physical properties

The physical properties of activated carbon play a crucial role in determining the

performance of supercapacitors, as evidenced by numerous prior studies [40,41]. Analyzing

key physical properties such as specific surface area, pore volume, and pore diameter is

essential for evaluating the suitability of activated carbon for supercapacitor electrodes. To this

end, we conducted N2 adsorption–desorption isotherm analysis of the synthesized activated

carbons (S_PACXs and T_PACYs), and the results are presented in Fig. 7 and Table 1.

Both the single-step and two-step processes exhibit a similar pattern where the specific

surface area of the resulting carbon materials initially increases and then decreases with rising

reaction temperature. This trend aligns with previous research findings, which attributed the

decrease in specific surface area at higher temperatures to pore blockage and structure collapse

induced by excessive activation [37]. Although the overall trends are comparable, notable

differences in the characteristics of each material appear based on the synthesis method. As

depicted in Table 1 and Fig. 7, the single-step method consistently yields activated carbons

with higher specific surface areas compared to the two-step method, particularly evident at 700

and 800°C. Notably, while activated carbons synthesized via the two-step process at these

temperatures exhibit specific surface areas smaller than that of commercial activated carbon

MSP20 (approximately 2300 m2 g–1), S_PAC700 and S_PAC800 demonstrate specific surface

areas of 3549 m2 g–1 and 3310 m2 g–1, respectively, surpassing those of the commercial

counterpart. The primary reasons for this enhanced specific surface area of activated carbons

produced via the single-step method will be further discussed in subsequent sections.

In supercapacitors, electrolyte ion adsorption primarily occurs within micropores, while

mesopores facilitate the smooth movement of electrolyte ions to the adsorption sites. The pore

structure characteristics of the synthesized activated carbons are illustrated in Fig. 7 and Table
1. It is observed that T_PACYs predominantly feature micropores, whereas S_PACXs exhibit

well-developed micropores and mesopores. This distinction suggests that S_PACXs are poised

to offer enhanced performance in terms of electrolyte ion movement and adsorption, owing to

their abundant micropores and mesopores. Specifically, with their high specific surface areas

and plentiful micropores and mesopores, S_PAC700 and S_PAC800 hold promise as superior

electrode materials for organic supercapacitors.

In the field of carbon materials production, the single-step method can be considered a

promising approach for enhancing the physical characteristics of activated carbons. Activated

carbons synthesized via the single-step method exhibit a high specific surface area along with

a rich formation of micropores and mesopores. This phenomenon could largely be attributed to

the unique characteristics of the single-step method, wherein carbonization and activation

occur concurrently. As shown in the proposed single-step mechanism in Fig. 5, KOH permeates

the PET particles before the initiation of the activation reaction within the temperature range

of 400–600°C. Subsequently, at temperatures exceeding 600°C, intensive activation reactions

proceed simultaneously from both the interior and exterior, leading to the increase of surface

area and the generation of abundant micropores and mesopores. Such a process enhances the

physical properties necessary for electrode materials, thus contributing to the anticipated

enhancement in supercapacitor performance.

By utilizing the single-step method, we could produce activated carbons with superior

properties as electrode materials. This efficient approach not only enhances time efficiency and

reduces energy consumption during synthesis but also yields a greater quantity of final product.

Ultimately, the single-step method is believed to be a highly effective strategy for activated

carbon production.
Fig. 7. (a) N2 adsorption–desorption isotherms of T_PACYs, (b) pore size distribution of T_PACYs, (c) N2

adsorption–desorption isotherms of S_PACXs and (d) pore size distribution of S_PACXs.


Table. 1. Physical properties of PET-derived activated carbons.

T_PAC600 T_PAC700 T_PAC800 T_PAC900 S_PAC600 S_PAC700 S_PAC800 S_PAC900

SBETa
1488.4 1983.8 2287.8 2076.7 1977.2 3549.0 3310.6 2467.7
(m2 g−1)

Davgb
1.9 1.9 2.0 2.4 1.8 2.0 2.2 2.5
(nm)

Vmicroc
0.8 1.0 0.9 0.9 1.0 1.4 1.0 0.6
(cm3 g−1)

Vmesod
0.0 0.0 0.2 0.4 0.0 0.5 1.1 1.0
(cm3 g−1)

Vtotale
0.8 1.0 1.1 1.3 1.0 1.9 2.1 1.6
(cm3 g−1)

a
Specific surface area
b
Average pore diameter
c
Micropore volume
d
Mesopore volume
e
Total pore volume
3.4. Crystallinity

The influence of activated carbon crystallinity on supercapacitor performance is well

established [42,43]. Typically, the crystallinity of carbon materials varies with heat treatment

temperature and duration. In this regard, concerns could arise regarding the lower crystallinity

of carbon materials produced by the single-step process due to reduced heat treatment time. To

analyze the crystallinity of PET-derived activated carbons, Raman spectroscopy was employed,

and the obtained Raman spectra along with the intensity ratio of the D peak (≈1350 cm−1) to

the G peak (≈1580 cm−1) are presented in Fig. 8.

Carbon materials form aromatic ring structures, resulting in a D peak (breathing vibration

of the ring) and a G peak (stretching vibration of the graphitic part, sp2 carbon), in Raman

spectra [44–46]. The intensity ratio of the D peak to the G peak obtained through Raman

spectroscopy is used to analyze the crystallinity of carbon materials. For amorphous carbon

materials like activated carbon, the ID/IG value directly correlates with crystallinity, according

to the Ferrari–Robertson relation. Conversely, in graphitic materials where a 2D peak is

observed, crystallinity is inversely proportional to the ID/IG value, aligning with the Tuinstra–

Koenig relation.

Irrespective of the manufacturing method, the ID/IG value increased as the temperature

escalated from 600 to 800°C. Additionally, a 2D peak, indicative of a significant increase in

crystallinity, was detected in activated carbon heat-treated at 900°C. These findings suggest

that heat treatment temperature influences the crystallinity of carbon materials, with a sharp

increase beyond a certain temperature threshold. We investigate crystallinity differences based

on the synthesis method by comparing S_PAC800 and T_PAC600, which exhibit similar ID/IG.

This suggests that to achieve equivalent crystallinity, a higher temperature is required in the

single-step method to compensate for reduced heat treatment time. While this presents a
drawback in terms of crystallinity for carbon materials produced by the single-step method, it

is essential to consider the overall advantages of the single-step method. The relatively lower

crystallinity can be offset by a significantly high specific surface area and abundant mesopores.

Considering the higher yield and efficient resource utilization, we anticipate that activated

carbons produced via the single-step method hold considerable promise as electrode materials.

Consequently, our next step involves assembling a coin-type organic supercapacitor using PET-

derived carbon materials and evaluating its electrochemical performance.

Fig. 8. Raman spectra of PET-derived activated carbons.

3.5. Electrochemical performance

Activated carbons derived from PET using the single-step method exhibit favorable

characteristics such as a high specific surface area, well-defined pore structure, and satisfactory

crystallinity. These attributes position them as promising candidates for high-performance


supercapacitor electrode materials. To assess their electrochemical performance, we

constructed coin-type organic supercapacitors using the S_PACXs produced through the

single-step method. For comparison, we also analyzed electrochemical properties under

identical conditions using T_PAC900, recognized for its outstanding performance among

activated carbons produced via the two-step method, and commercial activated carbon.

In Fig. 9, we present comprehensive data on cyclic voltammetry (CV), charge–discharge

(C–D), electrochemical impedance spectroscopy (EIS), and long-term durability for

supercapacitors fabricated with these activated carbons. Gravimetric capacitances of the

supercapacitors obtained from cyclic voltammograms were calculated using Eq. (1) to provide

a quantitative comparison of the electrochemical analysis results, as summarized in Table 2.

𝑎 𝐼 +|𝐼 |
𝑐
𝐶𝑔 = 2𝑚⋅(𝑑𝑉/𝑑𝑡) (1)

Here, in Eq. (1), Cg stands for gravimetric specific capacitance, Ia denotes the anodic current,

Ic represents the cathodic current, m denotes the weight of the active material loaded on both

electrodes, dV/dt signifies the scan rate.

At a slow scan rate of 5 mV s−1, all supercapacitors maintain a rectangular shape,

indicative of their stable performance. Notably, excluding T_PAC900 and S_PAC900, all

supercapacitors exhibit higher capacitance than MSP20, as evident from Fig. 9 and Table 2.

However, at a scan rate of 300 mV s−1, all supercapacitors exhibit a transformation into a rugby-

ball shape, signifying a loss in capacitance. Interestingly, S_PAC800 exhibits the highest initial

capacity of 40.0 F g−1, while S_PAC700, despite having a smaller initial capacity (34.4 F g−1),

demonstrates higher capacitance at a scan rate of 300 mV s−1. Moreover, S_PAC700 maintains

its initial capacity better under fast scan rate conditions compared to S_PAC900, attributed to

its higher crystallinity. Our findings diverge somewhat from the conventional understanding

that the capacitance of supercapacitors is primarily influenced by the specific surface area at
slow scan rates and electrical conductivity at fast scan rates [47,48]. This suggests that

properties other than specific surface area and electrical conductivity, such as mesopores, may

impact performance. We hypothesize that appropriate mesopores facilitate the movement of

electrolyte ions, enhancing capacitance retention even at fast scan rates. This hypothesis directs

Gravimetric capacitance (F g−1)

Scan rate (mV s−1) 5 10 30 50 100 300

MSP20 31.0 30.0 27.2 24.8 19.8 9.5

T_PAC900 20.8 19.9 18.4 17.6 15.9 11.4

S_PAC600 33.4 30.8 24.9 20.8 13.6 4.6

S_PAC700 34.4 33.3 31.8 30.2 29.5 24.4

S_PAC800 40.0 38.6 36.0 34.1 30.4 19.1

S_PAC900 21.3 20.6 19.3 18.7 17.3 12.8

our focus toward further electrochemical analyses.

Table 2. Specific capacitances of supercapacitors using MSP20 and PET-derived activated carbons calculated
from cyclic voltammograms.
Fig. 9. (a) Cyclic voltammograms at scan rate of 5 mVs-1, (b) cyclic voltammograms at scan rate of 300 mVs-1, (c) gravimetric capacitances, (d) electrochemical impedance
spectra and (e) long-term durability of supercapacitor using MSP20 and PET-derived activated carbons.
C–D analysis was conducted across a range of current densities (0.1, 0.5, 1, 3, 5, and 7 A

g-1) to evaluate the capacitance characteristics, as illustrated in Fig. 9 (c). Specific capacitance

values, both gravimetric (Cg) and volumetric (Cv), were calculated using Eq. (2) and Eq. (3),

respectively, and the results are listed in Table 3.

𝐼∙∆𝑡
𝐶𝑔 = 𝑚∙∆𝑉 (2)

𝐶𝑣 = 𝐶𝑔 × 𝜌𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑑𝑒 (3)

Here, in Eq. (2) and Eq. (3), I represents the discharge current, Δt stands for the discharge time,

m denotes the weight of the active material loaded on both electrodes, ΔV signifies the change

in voltage during the discharge process, and ρelectrode is determined by the ratio of the active

material loaded mass to the electrode volume.

The retention ratio (Rret) shown in Table 3 was calculated as the ratio of the specific

capacitance measured at a current density of 0.1 A g−1 to that measured at 7 A g−1. Gravimetric

capacitance results obtained through C–D analysis showed a trend similar to the results from

CV analysis. Specifically, under the condition of the lowest current density (0.1 A g−1), all

supercapacitors, excluding T_PAC900 and S_PAC900, exhibited higher capacitance compared

to the commercial activated carbon (MSP20). This difference may be attributed to the collapse

of the internal structure of carbon particles during high-temperature (900°C) heat treatment,

hindering effective utilization of the specific surface area and resulting in lower capacitance.

Consistent with the earlier CV analysis, S_PAC700 demonstrated the best maintenance of

initial capacity. Compared to MSP20, which had a Rret of 49.7%, S_PAC700 showed a

substantially higher Rret of 73.7%. The superior performance of S_PAC700 is believed to come

from its high specific surface area, decent crystallinity, and appropriate formation of mesopores.

Furthermore, all supercapacitors with carbon electrodes prepared using the single-step method,

except for S_PAC600, exhibited higher initial capacity and capacity retention than T_PAC900,
which performed best in the two-step process. This reinforces the superiority of activated

carbons produced through the single-step process in terms of electrochemical characteristics

for supercapacitor electrode materials.

To assess the long-term durability of the fabricated supercapacitors, C–D analysis was

repeated for 3000 cycles at a current density of 1 A g−1, with the results shown in Fig. 9(e).

After 3000 cycles, MSP20 retained only 86.0% of its capacity. However, all supercapacitors,

with the exception of S_PAC600, demonstrated superior capacity retention compared to

MSP20, indicating excellent long-term durability. Notably, S_PAC700 exhibited the highest

capacity retention, with an impressive value of 93.2%. Consequently, the electrochemical

characteristics analysis conducted through both CV and C–D confirmed that PET-derived

activated carbons synthesized using the single-step process outperformed those produced by

the two-step process, with S_PAC700 demonstrating the most superior electrochemical

performance.
Table 3. Specific capacitances of supercapacitors using MSP20 and PET-derived activated carbons.

Gravimetric capacitance [F g−1] Volumetric capacitance [F cm−3]

Current density Rreta


0.1 0.5 1 3 5 7 0.1 0.5 1 3 5 7
(A g−1) (%)

MSP20 30.5 28.8 26.9 20.1 14.2 10.1 24.0 22.7 21.2 15.8 11.2 7.5 32.8

T_PAC900 19.5 18.6 17.7 14.8 12.2 9.7 17.8 17.0 16.1 13.5 11.1 8.9 49.5

S_PAC600 32.3 28.7 24.9 13.2 5.6 1.6 26.1 23.2 20.2 10.7 4.5 1.3 5.0

S_PAC700 36.5 35.2 34.3 31.5 29.1 26.9 29.5 28.5 27.8 25.6 23.6 21.7 73.7

S_PAC800 39.5 37.7 36.4 32.3 28.5 25.1 24.5 23.4 22.6 20.1 17.7 15.7 63.5

S_PAC900 20.4 19.6 18.9 16.3 13.9 11.7 17.5 16.8 16.2 14.0 11.9 10.0 57.3
a
Retention ratio with increasing C–D: calculated from the ratio of specific capacitances at 0.1 and 7 A g−1.
Table 4. Sheet resistance of MSP20 and PET-derived activated carbons.

MSP20 T_PAC900 S_PAC600 S_PAC700 S_PAC800 S_PAC900

Rsheeta
716 249 1013 196 389 165
(Ohm sq.−1)
a
Sheet resistance

The electrochemical impedance spectroscopy (EIS) analysis was conducted to investigate

the electrical resistance characteristics of the assembled supercapacitors, with results depicted

as a Nyquist plot in Fig.10(d). This plot is divided into three distinct parts: bulk-solution

resistance, charge-transfer resistance, and Warburg resistance [49,50]. The bulk-solution

resistance, discerned by the x-intercept, mirrors the resistance originating from electrolyte ions.

Meanwhile, the charge-transfer resistance (Rct) displays as a semicircle, influenced by

electrolyte movement within electrode pores and resistance between the electrode and collector.

A small semicircle in the Rct region denotes the porous nature of the carbon material, signifying

its efficiency as a supercapacitor electrode. Thus, Rct plays a pivotal role in determining the

resistance inherent to the supercapacitor electrode material. The Warburg resistance, appearing

as an almost straight line after the Rct region, results from ion transfer limitations or irregular

paths.

The largest semicircle in the Rct region was observed in S_PAC600, correlating with its

notably low-capacity retention rate. Conversely, all other supercapacitors, excluding

S_PAC600, exhibited smaller semicircles than MSP20. Notably, S_PAC700 with the highest

capacity retention rate showed the smallest semicircle. Additionally, to verify the resistance

characteristics of the electrode material, frequency-dependent permittivity (FPP) analysis was

conducted employing OHP film as an insulator, with results presented in Table 4. The FPP

analysis outcomes mirrored trends observed in prior electrochemical analyses. S_PAC700,


performing well in CV and C–D analyses, exhibited the lowest resistance characteristic of 196

Ohm sq.−1, consistent with the EIS results.

In conclusion, PET-derived activated carbons produced via the single-step method

demonstrated outstanding electrochemical performance as high-performance supercapacitor

electrode materials. They are expected to outperform not only activated carbons from the two-

step method, but also commercial variants. Moreover, the single-step process proves

advantages in PET-derived activated carbon production, enhancing yield, shortening

production time, and reducing energy consumption.

4. Conclusions
This study presents a successful approach to producing high-quality activated carbons

from polyethylene terephthalate (PET), a commonly used plastic material, through a single-

step method. This method, which integrates carbonization and activation processes

simultaneously, offers a more efficient alternative to the conventional two-step method. The

single-step process effectively converts pure PET into carbon material, achieving a yield that

surpasses traditional methods. The mechanism of this process was analyzed in detail, revealing

three critical temperature-dependent stages that contribute to the high yield. The formation of

a melted potassium hydroxide (KOH) layer on the PET surface and its subsequent penetration

into the material were identified as key factors. The KOH layer creates favorable conditions

for PET conversion into carbon material, significantly enhancing the yield. Moreover, the

penetrated KOH leads to the development of a higher specific surface area and more extensive

mesopores compared to conventional methods. Particularly, the activated carbon designated as

S_PAC700 exhibited an exceptional specific surface area of 3549.0 m2 g−1, accompanied by a

well-developed pore structure and decent crystallinity. These features indicate superior
characteristics compared to carbons produced via the two-step method. Coin-type organic

supercapacitors were fabricated using PET-derived activated carbons and commercial

alternatives, followed by comprehensive electrochemical performance evaluations. Notably,

S_PAC700 exhibited remarkable initial capacitance (36.5 F g−1), high-capacity retention rate

(73.7%), low resistance characteristics (196 Ohm sq.−1), and excellent long-term durability,

outperforming T_PAC900 from the two-step process and surpassing commercial activated

carbon. In conclusion, this study demonstrates the effectiveness of the single-step heat

treatment method in upcycling PET into activated carbon for supercapacitors, offering

advantages in terms of time and energy efficiency compared to the two-step process while

yielding superior performance.


Acknowledgements

This work was supported by the National Research Foundation of Korea (NRF) grant

funded by the Korea government (MSIT) (2021R1F1A1046272).


References

[1] M. K. Hasan, M. Mahmud, A. K. A. A. Habib, S. M. A. Motakabber, S. Islam, J. Energy


Storage 41 (2021) 102940, https://doi.org/10.1016/j.est.2021.102940.

[2] M. A. Hasan, M. M. Hoque, A. Modamed, A. Ayob, Renew. Sustain. Energy Rev. 69 (2017)
771-789, http://dx.doi.org/10.1016/j.rser.2016.11.171.

[3] F. Ahmad, M. Khalid, B. K. Panigrahi, J. Energy Storage 43 (2021) 103153,


https://doi.org/10.1016/j.est.2021.103153.

[4] M. Aneke, M. Wang, Applied Energy 179 (2016) 350-337,


http://dx.doi.org/10.1016/j.apenergy.2016.06.097.

[5] L. Kouchachvil, W. Yaïci, E. Entchev, J. Power Sources 374 (2018) 237-248,


https://doi.org/10.1016/j.jporwsour.2017.11.040.

[6] B. K. Saikiaa, S. M. Benoy, M. Bora, J. Tamuly, M. Pandey, D. Bhattacharya, Fuel 282


(2020) 118796, https://doi.org/10.1016/j.fuel.2020.118796.

[7] A. González, E. Goikolea, J. A. Barrena, R. Mysyk, Renew. Sustain. Energy Rev. 58 (2016)
1189-1206, http://dx.doi.org/10.1016/j.rser.2015.12.249.

[8] Elzbieta Frackowiak, Q. Abbas, F. Béguin, J. Energy Chem. 22 (2013) 226-240,


https://doi.org/10.1016/S2095-4956(13)60028-5.

[9] L. Zhang, X. Hu, Z. Wang, F. Sun, D. G. Dorrell, Renew. Sustain. Energy Rev. 81 (2018)
1868-1878, http://dx.doi.org/10.1016/j.rser.2017.05.283.

[10] L. L. Zhang, X. S. Zhao, Chem. Soc. Rev. 38 (2009) 2520-2531,


https://doi.org/10.1039/B813846J.

[11] Z. Lin, E. Goikolea, A. Balducci, K. Naoi, P. L. Taberna, M. Salanne, G. Yushin, P. Simon,


Mater. Today 21 (4) (2018) 419-436, https://doi.org/10.1016/j.mattod.2018.01.035.

[12] O. Veneri, C. Capasso, S. Patalano, Appl. Energy 227 (2018) 312- 323,
https://doi.org/10.1016/j.apenergy.2017.08.086.

[13] D. Meena, R. Kumar, S. Gupta, O. Khan, D. Gupta, M. Singh, J. Energy Storage 72 (2023)
109323, https://doi.org/10.1016/j.est.2023.109323.

[14] R. T. Yadlapalli, R.K. R. Alla, R. Kandipati, A. Kotapati, J. Energy Storage 49 (2022)


104194, https://doi.org/10.1016/j.est.2022.104194.

[15] Poonam, K. Sharma, A. Arora, S.K. Tripathi, J. Energy Storage 21 (2019) 801-825,
https://doi.org/10.1016/j.est.2019.01.010.

[16] A. R. Selvaraj, A. Muthusamy, I.-Cho, H.-J. Kim, K. Senthil, K. Prabakar, Carbon 174
(2021) 463-474, https://doi.org/10.1016/j.carbon.2020.12.052.

[17] A. Afif, S. M. Rahman, A. T. Azad, J. Zaini, M. A. Islan, A. K. Azad, J. Energy Storage 25


(2019) 100852, https://doi.org/10.1016/j.est.2019.100852.

[18] K. K. Patel, T. Singhal, V. Pandey, T. P. Sumangala, M. S. Sreekanth, J. Energy Storage


44 (2021) 103366, https://doi.org/10.1016/j.est.2021.103366.

[19] S. P. M, D. P, R. R, Sustain. Chem. Pharm. 16 (2020) 100243,


https://doi.org/10.1016/j.scp.2020.100243.

[20] I. Yang, D. Kwon, J. Yoo, M.-S. Kim, J. C. Jung, Curr. Appl. Phys. 19 (2019) 89-96,
https://doi.org/10.1016/j.cap.2018.10.010.

[21] I. Yang, D. Kwon, M.-S. Kim, J. C. Jung, Carbon 132 (2018) 503-511,
https://doi.org/10.1016/j.carbon.2018.02.076.

[22] N. Talreja, S. H. Jung, L. T. H. Yen, T. Y. Kim, Chem. Eng. J. 379 (2020) 122332,
https://doi.org/10.1016/j.cej.2019.122332.

[23] I. Yang, S.-G. Kim, S. H. Kwon, M.-S. Kim, J. C. Jung, Electrochem. Acta 223 (2017) 21-
30, http://dx.doi.org/10.1016/j.electacta.2016.11.177.

[24] S. Chen, Z. Liu, S. Jiang, H. Hou, Sci. Total Environ. 710 (2020) 136250,
https://doi.org/10.1016/j.scitotenv.2019.136250.

[25] O. S. Alimi, J. F. Budarz, L. M. Hernandez, N. Tufenkji, Environ. Sci. Technol. 52 (2018)


1704-1724, https://10.1021/acs.est.7b0555.

[26] S. Kubowicz, A. M. Booth, Environ. Sci. Technol. 51 (2017) 12058-12060,


https://doi.org/10.1021/acs.est.7b04051.

[27] H. Zhang, X.-L. Zhou, L.-M. Shao, F. Lü, P.-J. He, Sci. Total Environ. 772 (2021) 145309,
https://doi.org/10.1016/j.scitotenv.2021.145309.

[28] S. A. Bhat, V. Kumar, S. Kumar, A.E. Atabani, I. A. Badruddin, K.-J. Chae, Fuel 337 (2023)
127125, https://doi.org/10.1016/j.fuel.2022.127125.

[29] O. Awogbemi, D. V. V. Kallon, J. Energy Inst. 106 (2023) 101154,


https://doi.org/10.1016/j.joei.2022.101154.

[30] S. Sharifian, N. A.-Kolur, J. Anal. Appl. Pyrolysis 163 (2022) 105496,


https://doi.org/10.1016/j.jaap.2022.105496.

[31] X. N. Wei, T. T. Li, ACS Omega 6 (2021) 5607-5618,


https://dx.doi.org/10.1021/acsomega.0c06032.

[32] Y. Gao, Q. Yue, B. Gao, A. Li, Sci. Total Environ. 746 (2020) 141094,
https://doi.org/10.1016/j.scitotenv.2020.141094.

[33] N. Z. M. Azmi, A. Buthiyappan, A. A. A. Raman, M. F. A. Patah, S. Sufian, J. Ind. Eng.


Chem. 116 (2022) 1-20, https://doi.org/10.1016/j.jiec.2022.08.021.

[34] W. Chen, X. Wang, C. Liu, M. Luo, P. Yang, X. Zhou, Waste Manag. 102 (2020) 330-339,
https://doi.org/10.1016/j.wasman.2019.10.058.
[35] P. Ozpinar, C. Dogan, H. Demiral, U. Morali, S. Erol, C. Samdan, D. Yildiz, I. Demiral,
Renew. Energy 189 (2022) 535-548, httpsr://doi.org/10.1016/j.renene.2022.02.126.

[36] S. Brunauer, P. H. Emmett, E. Teller, J. Am. Chem. Soc. 60 (2) (1938) 309-319,
https://doi.org/10.1021/ja01269a023.

[37] I. Yang, M. Jung, M.-S. Kim, D. Choi, J. C. Jung, J. Mater. Chem. A 9 (2021) 9815-9825,
https://doi.org/10.1039/d1ta00765c.

[38] L. Fan, P. Sun, L. Yang, Z. Xu, J. Han, Korean J. Chem. Eng. 37 (1) (2020) 166-175,
https://doi.org/10.1007/s11814-019-0414-8.

[39] J. W.F. Chia, O. Sawai, T. Nunoura, Waste Manag. 108 (2020) 62-69,
https://doi.org/10.1016/j.wasman.2020.04.035.

[40] X. Liu, Y. Wen, X. Chen, T. Tang, E. Mijowsk, Sci. Total Environ. 723 (2020) 138055,
https://doi.org/10.1016/j.scitotenv.2020.138055.

[41] M. Jung, I. Yang, J. Yoo, M.-S. Kim, J. C. Jung, J. Electrochem. Soc. 168 (2021) 080532,
https://doi.org/10.1149/1945-7111/ac13d3.

[42] I. Yang, J. H. Mok, M. Jung, J. Yoo, M.-S. Kim, D. Choi, J. C. Jung, Macromol. Rapid
Commun. 42 (2022) 220006, https://doi.org/10.1002/marc.202200006.

[43] M. Jung, I. Yang, D. Choi, J. Lee, J. C. Jung, Korean J. Chem. Eng. 40 (10) (2023) 2442-
2454, https://doi.org/10.1007/s11814-023-1466-3.

[44] A. C. Ferrari, J. Robertson Phys. Rev. B 61 (20) (2000) 14095,


https://doi.org/10.1103/PhysRevB.61.14095.

[45] D. Choi, D. Jang, H.-I. Joh, E. Reichmanis, S. Lee, Chem. Mater. 29 (2017) 9518-9527,
https://doi.org/10.1021/acs.chemmater.7b03737.

[46] F. TUINSTRA, J. L. KOENIG, J. Chem. Phys. 53 (3) (1970) 1126-1130,


https://doi.org/10.1063/1.1674108.

[47] Y. Lu, S. Zhang, J. Yin, C. Bai, J. Zhang, Y. Li, Y. Yang, Z. Ge, M. Zhang, L. Wei, M. Ma,
Y. Ma, Y. Chen, Carbon 124 (2017) 64-71, http://dx.doi.org/10.1016/j.carbon.2017.08.044.

[48] F. Sun, D. Wu, J. Gao, T. Peia, Y. Chen, K. Wang, H. Yang, G. Zhao, J. Power Sources 477
(2020) 228759, https://doi.org/10.1016/j.jpowsour.2020.228759.

[49] H. D. Yoo, J. H. Jang, J. H. Ryu, Y. Park, S. M. Oh, J. Power Sources 267 (2014) 411-420,
http://dx.doi.org/10.1016/j.jpowsour.2014.05.058.

[50] B.-A. Mei, O. Munteshari, J. Lau, B. Dunn, L. Pilon, J. Phys. Chem. C 122 (2018) 194-
206, http://dx.doi.org/10.1021/acs.jpcc.7b10582.
Figure Click here to access/download;Figure;Figure.pptx

Fig.1. Scheme of PET-derived activated carbons fabrication process.


Fig. 2. Scheme of the fabrication process for supercapacitor.
Fig. 3. TGA date of PET obtained under nitrogen condition.
Fig. 4. Yield of PET-derived activated carbons.
Fig. 5. Scheme of the single-step reaction process.
Fig. 6. SEM images of (a) PET, (b) S_PAC700, and (c) T_PAC700.
Fig. 7. (a) N2 adsorption–desorption isotherms of T_PACYs, (b) pore size distribution of T_PACYs, (c) N2 adsorption–desorption isotherms of
S_PACXs and (d) pore size distribution of S_PACXs.
Fig. 8. Raman spectra of PET-derived activated carbons.
Fig. 9. (a) Cyclic voltammograms at scan rate of 5 mVs-1, (b) cyclic voltammograms at scan rate of 300 mVs-1, (c) gravimetric capacitances, (d)
electrochemical impedance spectra and (e) long-term durability of supercapacitor using MSP20 and PET-derived activated carbons.
Table (Editable version) Click here to access/download;Table (Editable version);Table.docx

Tables

Table 1. Physical properties of PET-derived activated carbons.

T_PAC600 T_PAC700 T_PAC800 T_PAC900 S_PAC600 S_PAC700 S_PAC800 S_PAC900

SBETa
1488.4 1983.8 2287.8 2076.7 1977.2 3549.0 3310.6 2467.7
(m2 g–1)

Davgb
1.9 1.9 2.0 2.4 1.8 2.0 2.2 2.5
(nm)

Vmicroc
0.8 1.0 0.9 0.9 1.0 1.4 1.0 0.6
(cm3 g–1)

Vmesod
0.0 0.0 0.2 0.4 0.0 0.5 1.1 1.0
(cm3 g–1)

Vtotale
0.8 1.0 1.1 1.3 1.0 1.9 2.1 1.6
(cm3 g–1)

a
Specific surface area
b
Average pore diameter
c
Micropore volume
d
Mesopore volume
e
Total pore volume
Table 2. Specific capacitances of supercapacitors using MSP20 and PET-derived activated carbons calculated from cyclic voltammograms.

Gravimetric capacitance (F g–1)

Scan rate (mV s–1) 5 10 30 50 100 300

MSP20 31.0 30.0 27.2 24.8 19.8 9.5

T_PAC900 20.8 19.9 18.4 17.6 15.9 11.4

S_PAC600 33.4 30.8 24.9 20.8 13.6 4.6

S_PAC700 34.4 33.3 31.8 30.2 29.5 24.4

S_PAC800 40.0 38.6 36.0 34.1 30.4 19.1

S_PAC900 21.3 20.6 19.3 18.7 17.3 12.8


Table 3. Specific capacitances of supercapacitors using MSP20 and PET-derived activated carbons.

Gravimetric capacitance (F g–1) Volumetric capacitance (F cm–3)

Current density Rreta


0.1 0.5 1 3 5 7 0.1 0.5 1 3 5 7
(A g–1) (%)

MSP20 30.5 28.8 26.9 20.1 14.2 10.1 24.0 22.7 21.2 15.8 11.2 7.5 32.8

T_PAC900 19.5 18.6 17.7 14.8 12.2 9.7 17.8 17.0 16.1 13.5 11.1 8.9 49.5

S_PAC600 32.3 28.7 24.9 13.2 5.6 1.6 26.1 23.2 20.2 10.7 4.5 1.3 5.0

S_PAC700 36.5 35.2 34.3 31.5 29.1 26.9 29.5 28.5 27.8 25.6 23.6 21.7 73.7

S_PAC800 39.5 37.7 36.4 32.3 28.5 25.1 24.5 23.4 22.6 20.1 17.7 15.7 63.5

S_PAC900 20.4 19.6 18.9 16.3 13.9 11.7 17.5 16.8 16.2 14.0 11.9 10.0 57.3
a
Retention ratio with increasing C–D: calculated from the ratio of specific capacitances at 0.1 and 7 A g–1.
Table 4. Specific capacitances of supercapacitors using YP50f and PETKs calculated from cyclic voltammograms.

MSP20 T_PAC900 S_PAC600 S_PAC700 S_PAC800 S_PAC900

Rsheeta
716 249 1013 196 389 165
(Ohm sq.-1)

a
Sheet resistance

You might also like