Ref 1 RD

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

Accepted Manuscript

Title: Development of a cost-effective CO2 adsorbent from


petroleum coke via KOH activation

Authors: Eunji Jang, Seung Wan Choi, Seok-Min Hong,


Sangcheol Shin, Ki Bong Lee

PII: S0169-4332(17)32408-X
DOI: http://dx.doi.org/10.1016/j.apsusc.2017.08.075
Reference: APSUSC 36912

To appear in: APSUSC

Received date: 3-3-2017


Revised date: 27-7-2017
Accepted date: 9-8-2017

Please cite this article as: Eunji Jang, Seung Wan Choi, Seok-Min Hong,
Sangcheol Shin, Ki Bong Lee, Development of a cost-effective CO2
adsorbent from petroleum coke via KOH activation, Applied Surface
Sciencehttp://dx.doi.org/10.1016/j.apsusc.2017.08.075

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Development of a cost-effective CO2 adsorbent from

petroleum coke via KOH activation

Eunji Jang†, Seung Wan Choi†, Seok-Min Hong, Sangcheol Shin, Ki Bong Lee*

Department of Chemical and Biological Engineering, Korea University, 145 Anam-ro,

Seongbuk-gu, Seoul 02841, Republic of Korea

†These authors contributed equally to this work.

*Corresponding author

Tel.: +82 2 3290 4851; FAX: +82 2 926 6102; E-mail: kibonglee@korea.ac.kr (K.B. Lee)

1
Graphical abstract

KOH
activation

Petroleum coke Microporous carbon

Highlights

 Cost effective CO2 adsorbent could be prepared from petroleum coke.


 CO2 uptakes have good correlation with narrow micropores (< 0.8 nm).
 The prepared microporous carbon showed excellent CO2 adsorption performanc
e.

Abstract
The capture of CO2 via adsorption is considered an effective technology for decreasing global

warming issues; hence, adsorbents for CO2 capture have been actively developed. Taking into account

cost-effectiveness and environmental concerns, the development of CO2 adsorbents from waste

materials is attracting considerable attention. In this study, petroleum coke (PC), which is the carbon

residue remaining after heavy oil upgrading, was used to produce high-value-added porous carbon for

CO2 capture. Porous carbon materials were prepared by KOH activation using different weight ratios

of KOH/PC (1:1, 2:1, 3:1, and 4:1) and activation temperatures (600, 700, and 800 C). The specific

surface area and total pore volume of resulting porous carbon materials increased with KOH amount,

reaching up to 2433 m2/g and 1.11 cm3/g, respectively. The sample prepared under moderate conditions

with a KOH/PC weight ratio of 2:1 and activation temperature of 700 C exhibited the highest CO2

2
adsorption uptake of 3.68 mmol/g at 25 C and 1 bar. Interestingly, CO2 adsorption uptake was linearly

correlated with the volume of micropores less than 0.8 nm, indicating that narrow micropore volume is

crucial for CO2 adsorption. The prepared porous carbon materials also exhibited good selectivity for

CO2 over N2, rapid adsorption, facile regeneration, and stable adsorption–desorption cyclic performance,

demonstrating potential as a candidate for CO2 capture.

Keywords: CO2 adsorption; porous carbon; petroleum coke; KOH activation; narrow micropore

3
1. Introduction

Global warming and the related climate change issues, attributed to the increase of greenhouse gas

(GHG) emissions, have attracted significant attention worldwide [1]. In particular, CO2 is a major GHG,

and its emissions from the combustion of fossil fuels as well as industrial processes account for 78% of

the total increase of GHG emissions [2]. According to the International Energy Agency (IEA), CO2

emissions have significantly increased since the industrial revolution. Currently, the concentration of

CO2 in the atmosphere is greater than 400 ppm, which is ~40% greater than that in the mid-1800s, with

an average growth of 2 ppm per year in the last ten years [3]. For reducing the CO2 emissions from the

energy sector, many studies have been conducted on the development of CO2 capture and storage (CCS)

technology. As the CO2 capture step accounts for two-thirds of the total cost of CCS [4,5], lots of studies

have been conducted to improve capture technologies. Currently, absorption using an aqueous amine

solution is commercially utilized for CO2 capture [6]; however, this approach is limited by several

disadvantages, such as the corrosion of process equipment and high consumption of energy. To

overcome these drawbacks, alternative technologies need to be developed. Membrane separation is one

of potential techniques for CO2 capture because of the advantages of high compactness and continuous

operation. However, the commercialization of membrane separation was hindered by low separation

efficiency at low feed concentration and low thermal stability in case of polymeric membrane and

vulnerability to cracking in case of inorganic or carbon membrane [7]. Recently, various studies have

been conducted to develop highly efficient membranes: e.g. thermally stable polymeric membrane

based on polybenzimidazole and mixed-ion conducting membrane for fast mass transport [8–10].

Adsorption using solid materials is promising as this approach exhibits advantages of low energy

requirement for regeneration, scale-up feasibility, and mild operating conditions. Among various

adsorbents, e.g., metal–organic frameworks [11,12], zeolites [13], activated carbon [14,15], and

mesoporous silica [16], activated carbon is attractive because of advantages, e.g., cost-effectiveness,

chemical and mechanical stabilities, and tunable pore structures.

4
Activated carbon can be obtained from several carbon-containing precursors (e.g., bones, phenolic

resins, polyacrylonitrile, coals, and lignocellulosic materials) [17–20]. Several studies have investigated

the development of CO2 adsorbents from inexpensive waste materials, which are stockpiled. The

utilization of these waste materials can not only resolve increasing environmental concerns associated

with increasing amounts of landfilled waste but also produce cost-effective CO2 adsorbents with

enhanced competitiveness. Plaza et al. have prepared CO2 adsorbents from a cost-effective, relatively

abundant agricultural byproduct such as almond shells [21]. Porous carbon was synthesized by the

carbonization of almond shells, followed by CO2 activation, affording carbon with a CO2 capture

capacity of 2.7 mmol/g at 25 C and 1 bar; this value is comparable to that of commercial activated

carbons. Fan et al. have fabricated nitrogen-doped microporous carbon from chitosan using K2CO3 as

the activating agent, and their cost-effective and environmentally friendly adsorbent exhibited a capture

capacity of 3.86 mmol/g at 25 C and 1 bar [22]. Parchetti et al. have produced cost-effective

carbonaceous adsorbents from the empty fruit bunch of oil palm trees; it is otherwise used as mulch or

incinerated in mills, resulting in air pollution [23]. The activated carbon prepared by hydrothermal

carbonization coupled with chemical activation exhibited a CO2 adsorption capacity of 3.71 mmol/g at

25 C and 1 atm.

Petroleum coke (PC) is another type of waste residue remaining after heavy oil upgrading. For

meeting the increasing demands for light oil, heavy oil fractions must be converted into light and more

valuable products via various upgrading technologies [24]. During upgrading pyrolysis, heavy (residual)

oil fractions are thermally cracked to produce light products, which are present in the gas product stream,

and a solid carbon-rich byproduct (or PC) is left behind. As PC contains high carbon and low volatile

and ash contents [25], it can be a good carbon precursor for activated carbons. By adequate activation

processes, industrial waste PC can be transformed into cost-effective value-added products. Recently,

Yang et al. reported porous nitrogen-doped carbon developed by combining ammoxidation with KOH

activation, and it showed high CO2 adsorption uptake and CO2/N2 selectivity [26].

5
In this study, porous carbon was synthesized from waste PC via KOH activation for use as an

adsorbent for CO2. PC was obtained as a byproduct during the pyrolysis upgrading of vacuum residue

(VR) on a laboratory scale. By varying the KOH to PC weight ratio and activation temperature, samples

with various surface areas and pore volumes were prepared. Then, the prepared porous carbon samples

were tested for CO2 adsorption. From the experimental results, the relation between porosities and CO2

adsorption uptakes were investigated and the pore size crucial for CO2 uptake was determined.

Furthermore, CO2/N2 selectivity, adsorption kinetics, and cyclic stability measurements were conducted

to examine the practical utility of the prepared adsorbents.

2. Experiment

2.1. Preparation of petroleum coke

PC used herein was obtained from the pyrolysis of VR on a laboratory scale. First, 50 g of VR

produced from a commercial vacuum distillation process (SK Innovation, Republic of Korea) was

added into a semi-batch reactor. Then, the reactor was heated to 450 °C and maintained at that

temperature for 2 h for pyrolysis. During pyrolysis, N2 was flowed through the reactor to discharge

volatile products. PC was produced as a residual solid product, which remained in the reactor after the

experiment. According to the results obtained from elemental analysis (Flash EA 1112 series, CE

Instruments) shown in Table 1, carbon is mainly present in VR and PC, the content of which increases

after pyrolysis.

2.2. Preparation of porous carbons

Several samples were prepared by mixing 2 g of PC with a solution containing different amounts of

KOH (KOH/PC weight ratios of 1:1, 2:1, 3:1, and 4:1) and stirring the resulting mixture at 60 C for 2

h. The mixture was then placed into an alumina boat and dried at 110 °C for 24 h. After drying, the

alumina boat was inserted into a horizontal quartz tube furnace and heated at a rate of 5 °C/min to an

activation temperature of 600, 700, and 800 °C under N2. Next, the furnace was maintained at the

6
activation temperature for 1 h, followed by cooling to room temperature. The resulting sample was

washed with a 10% HCl solution to remove the remaining potassium, followed by washing several

times with deionized water until the pH of the filtrate was neutral. Finally, the sample was dried at 110

C for 24 h. The obtained activated porous carbon was referred to as PC-x:y-z (where, x:y represents the

weight ratio of KOH to PC and z denotes the activation temperature).

2.3. Characterization

Scanning electron microscopy (SEM, Hitachi S-4800) and high-resolution transmission electron

microscopy (HRTEM, Tecnai G2 F30ST) were employed to investigate sample morphologies. N2

adsorption–desorption isotherms were measured at 77 K on a Micromeritics ASAP 2020 volumetric

adsorption analyzer to analyze the textural properties of porous carbon samples. Before each

measurement, samples were degassed under vacuum at 150 C for 12 h. The specific surface area was

calculated by the Brunauer–Emmett–Teller (BET) analysis from the adsorption data in the relative

pressure range of 0.01–0.1. The total pore volume was obtained from the adsorbed amount at a relative

pressure of 0.99, and the micropore volume was determined using the Dubinin–Radushkevich (D–R)

equation. Non-localized density functional theory (NLDFT) was employed to evaluate the pore size

distribution from the N2 adsorption data assuming a slit-pore model. Using the same volumetric

adsorption analyzer, the CO2 adsorption isotherms were measured at three temperatures of 0, 25, and

50 C, respectively, up to 1 bar. Thermogravimetric analysis (TGA, Q50, TA instruments) was used for

investigating the thermal decomposition of PC mixed with KOH. TGA was also employed for kinetic

and cyclic stability measurements of the adsorption of CO2 on porous carbon. The chemical surface of

samples was analyzed using Fourier transform infrared spectroscopy (FT-IR, LabRam Aramis IR2,

Horiba) and X-ray photoelectron spectroscopy (XPS, K-alpha, Thermo).

3. Results and discussion

3.1. Characteristics of KOH-activated PCs

7
Fig. 1 shows the SEM images of the morphologies of pristine PC and activated PC. Pristine PC

exhibited a relatively smooth surface, while irregularly shaped activated PC exhibited a rough surface,

indicative of a reaction occurring between KOH and carbon atoms during activation. Depending on

activation conditions, activated PC exhibited different morphologies. In particular, PC-2:1-700 and PC-

3:1-700 exhibited many cavities as compared with those observed for PC-1:1-700 (Fig. 1b–d). During

activation under harsh conditions with a KOH/PC weight ratio of 4, small pieces of broken PC particles

with sharp edges were observed (Fig. 1e). The similar morphology characteristics were detected in the

samples activated at different temperatures. PC-2:1-600 showed relatively smoother edges than those

of samples activated at higher temperature, and PC-2:1-800 had similar morphology to that of PC-4:1-

700 because of harsh activation condition (Fig. 1f and g). As can be observed in the HRTEM image of

PC-2:1-700 (Fig. 1h), randomly distributed micropores were observed in the amorphous carbon

structure.

Fig. 2 shows the N2 adsorption–desorption isotherms of porous carbon samples prepared with

different KOH/PC weight ratios and activation temperatures. All samples exhibited type I isotherms

according to the IUPAC classification [27]. The isotherms exhibited significant N2 uptake at low

relative pressures of P/P0 < 0.1 and a plateau for P/P0 > 0.1, implying that micropores are predominantly

present in porous carbon samples. The adsorption–desorption curves were identical without any

hysteresis loop. The samples activated at a low KOH/PC weight ratio exhibited an abrupt knee in the

adsorption isotherm curves, indicative of narrow microporosity, and the samples activated at high

KOH/PC weight ratio presented a wider knee, attributed to broader microporosity (Fig. 2a). At a fixed

KOH/PC weight ratio of 2:1, the knee in the adsorption isotherm curves became wider as the activation

temperature increased (Fig. 2b). Table 2 summarizes the textural properties of porous carbon samples

estimated from the N2 adsorption isotherms at 77 K. With increasing KOH/PC weight ratio, the specific

surface area and total pore volume of samples significantly increased from 915 m2/g and 0.38 cm3/g to

2433 m2/g and 1.11 cm3/g, respectively. Large amounts of KOH resulted in high carbon consumption

and created many micropores according to the following reaction: 6KOH + 2C  2K + 3H2 + 2K2CO3

8
[28]. These results indicate that the pore structure of porous carbon samples significantly depends on

the KOH/PC weight ratio during activation. Activation temperature is also an important factor to control

the textural properties of the samples. At a fixed KOH/PC weight ratio of 2:1, textural properties became

improved with increasing activation temperature.

Fig. 3 shows the pore size distribution and cumulative pore volume as obtained from N2 adsorption

isotherms by NLDFT assuming a slit-pore model. With increasing KOH/PC weight ratio, the proportion

of large micropores increased (Fig. 3a). The samples activated under mild conditions (KOH/PC weight

ratios of 1:1, 2:1, and 3:1) were mainly composed of micropores centered at 0.6 and 0.8 nm. On the

other hand, the micropore size increased, and small mesopores were newly formed in PC-4:1-700. This

tendency was further confirmed from the curve of the cumulative pore volume (Fig. 3b). For the samples

activated with KOH/PC weight ratios of 1:1, 2:1, and 3:1, the cumulative pore volume increased with

increasing size of up to 1.5 nm and then reached a plateau. However, the cumulative pore volume

continuously increased in the microporous region for PC-4:1-700. At a fixed KOH/PC weight ratio of

2:1, micropores actively developed up to the activation temperature of 700 °C (PC-2:1-700) and larger

pores were detected at higher activation temperature than 700 °C (PC-2:1-800). For the cumulative pore

volume, PC-2:1-700 showed the highest value until the pore width of ~1.3 nm, however, PC-2:1-800

had the highest value at the pore width over ~1.3 nm (Fig. 3d). As activated PC comprises a large

volume of narrow micropores, it is expected to be an effective adsorbent for capturing CO2 [29, 30].

The thermal decomposition of PC mixed with KOH in the ratio of 2:1 was analyzed using TGA to

further understand the effect of activation temperature (Fig. 4). The weight decrease up to ~100 °C is

from the evaporation of moisture and the additional weight decrease until ~350 °C can be attributed to

pore development through the reaction with KOH. After that, no further weight decrease was detected

until ~600 °C. However, when temperature reached ~700 °C, the sample weight started to decrease due

to further development of pores resulting from the reaction with KOH. When the activation temperature

increased over 750 °C, a sharp decrease of sample weight was observed, relating to pore enlargement

9
and increasing surface area. Based on the thermal decomposition results, the optimum temperature for

the activation of PC was expected as 700 °C.

The chemical surface of pristine PC and PC-derived porous carbons was investigated using FT-IR

(Fig. 5). Nitrogen, sulfur, hydrogen, and oxygen were detected as heteroatoms in the porous carbons

and various bonds of heteroatoms to carbon were observed. The absorption bands at the wavenumber

of 1562, 1750, and 2928 cm–1 are attributed to stretching vibrations of C=C, C=O, and C–H, respectively

[31–33]. Nitrogen and sulfur exist in the form of C=N, C–S, C–N, and S–H and their absorption bands

are observed at the wavenumber of 1640 (C=N), 855 (C–S), 1195 (C–S), and 2343 cm–1 (C–N or S–

H), respectively [31,33,34]. However, no distinct changes of FT-IR spectra were detected with changing

KOH amount or activation temperature.

Also, the chemical surfaces were studied in detail by XPS (Fig. 6). In the S2p spectra (Fig. 6a), it can

be found that some of sulfur components originated from PC were oxidized after KOH activation.

Therefore, the PC-derived porous carbons have both neutral sulfur (163.6 and 164.7 eV) and oxidized

sulfur (168.0 eV), while PC contains only neutral sulfur [35]. Since sulfur has weak acidity, it is

anticipated that sulfur-doped carbon has low affinity to CO2. However, when sulfur exists in an oxidized

form (-SO or -SO2), the acidity of sulfur makes oxygen negatively charged. The negatively charged

oxygen has high affinity to CO2 [36,37]. In the N1s XPS spectra (Fig. 6b), a weak C–N peak centered

at binding energy of 400 eV could be seen [38].

3.2. Adsorption of CO2 on KOH-activated PCs

Fig. 7a and b show the CO2 adsorption isotherms at 25 C and up to 1 bar for samples activated with

different KOH/PC weight ratios and activation temperatures, respectively. Table 3 summarizes their

adsorption uptake values. At 25 C and 1 bar, PC-2:1-700 exhibited the highest CO2 uptake of 3.68

mmol/g, indicating that an optimal amount of KOH and activation temperature are required for the

activation of PC to attain the maximum CO2 uptake. Although specific surface area and total pore

volume continuously increased with increasing KOH/PC weight ratio and activation temperature, the

10
CO2 adsorption uptake reached the maximum at a KOH/PC weight ratio of 2 and activation temperature

of 700 C. This result indicates that CO2 uptake is not directly related to the specific surface area or

total pore volume (Fig. 7c). Pore size increased, as well as pores were generated, at high KOH/PC

weight ratios and activation temperature, resulting in the increase of surface area and total pore volume,

but a large pore size is not adequate for the pore filling mechanism for CO2 adsorption at 1 bar. Several

studies have reported that narrow micropores less than 0.8 nm or 1 nm mainly contribute to the CO 2

adsorption uptake [39–41]. To investigate the effect of the narrow micropore volume on CO2 adsorption,

Fig. 7d plots the relationship between the volumes of narrow micropores less than 0.8 nm and the

amount of CO2 adsorbed at 25 C and 1 bar. With increasing volume of micropores less than 0.8 nm,

the CO2 uptake increased, with a correlation coefficient (R2) of 0.96, indicating that narrow micropores

significantly affect the CO2 adsorption uptake at 25 C and 1 bar. This is because narrow micropores

exhibit strong adsorption potentials, which can increase the strength of interaction between CO 2

molecules and pore walls [42]. Hence, to achieve high CO2 adsorption uptake, it is imperative to

synthesize adsorbents containing narrow micropores (<0.8 nm) by adjusting activation conditions.

Interestingly, the CO2 uptake for PC-4:1-700 was greater than that observed for PC-1:1-700 at

pressures greater than 0.5 bar, but the CO2 uptake for PC-1:1-700 was greater than that observed for

PC-4:1-700 at low pressures of less than 0.5 bar. This result was attributed to the reduced amount of

very small pores around 0.6 nm in PC-4:1-700 as shown in Fig. 3a. Another reason can be the effect of

nitrogen, which is known to serve as a basic site for attracting acidic CO2 molecules. Although samples

were not intentionally doped with nitrogen, raw PC as obtained from pyrolysis still exhibited low

amounts of nitrogen (Table 1). The amount of nitrogen in the sample decreased after activation as

nitrogen was oxidized and removed by KOH. As PC-1:1-700 (0.63%) exhibited a relatively high

nitrogen content as compared with that observed for PC-4:1-700 (0.21%), PC-1:1-700 can adsorb more

CO2 at low pressures. The nitrogen content has been reported to be crucial for the CO 2 uptake at low

pressures [43,44].

11
Fig. 8 shows the effect of temperature on the CO2 adsorption uptake of PC-2:1-700 at 0, 25, and 50

C. Depending on the adsorption temperature, different CO2 uptake values were observed at the pressure

of 1 bar: 6.08 mmol/g at 0 °C and 2.14 mmol/g at 50 °C. Low CO2 uptake at high temperature indicates

that the adsorption of CO2 on KOH-activated PC is an exothermic process.

To investigate the adsorption of CO2 on porous carbon, CO2 adsorption isotherms of PC-2:1-700 at

0, 25, and 50 C were fitted with the Langmuir isotherm model (Fig. 8). The widely accepted Langmuir

isotherm model is a simple adsorption model, which is suitable for providing a thermodynamic basis to

describe adsorption equilibria [45]. The Langmuir isotherm model is expressed as follows:

qmax K L P
qe = (1)
1 + KL P

Here, qe is the adsorbed amount at equilibrium, qmax is the maximum adsorption capacity, KL is the

Langmuir constant, and P is the pressure. Table 4 summarizes the parameters obtained from the

Langmuir model fitting and correlation coefficients (R2). The CO2 adsorption isotherm data were well

described by the Langmuir model, with sufficiently high correlation coefficients.

As post-combustion flue gases comprise 15% of CO2, with the remainder being mainly N2 [46,47],

it is imperative to selectively capture CO2 over N2 for practical applications. To compare CO2 and N2

uptakes on PC-2:1-700, N2 adsorption isotherm was measured under the same conditions as those

measured for CO2 (25 C and up to 1 bar). As shown in Fig. 9a, the adsorption uptake for N2 was only

0.48 mmol/g at 25 C and 1 bar, which was significantly less than that of CO2 at any pressure. For

calculating the selectivity of CO2 over N2 in flue gas, the ideal adsorbed solution theory (IAST) model

was applied to predict the binary gas mixture equilibria using single-component isotherms [48–51]. The

IAST selectivity (SCO2/N2) was calculated as follows:

qCO2 / qN 2
SCO2 /N 2 = (2)
pCO2 / pN 2

Here, qCO2 and qN2 are the adsorbed amounts of CO2 and N2, respectively, obtained from the single-

component adsorption isotherm. pCO2 and pN2 are the partial pressures of CO2 (0.15 bar) and N2 (0.85
12
bar) in flue gas, respectively. At a total pressure of 1 bar and 25 C, the selectivity of PC-2:1-700 for

CO2 over N2 was 13.7, indicating that CO2 is preferentially separated effectively from N2 (Fig. 9b).

Fig. 10 shows the adsorption kinetics of CO2 on PC-2:1-700 at 30 C and 1 bar. Kinetic analysis was

conducted by TGA. First, the sample was pretreated under N2 at 150 C for 4 h to remove any pre-

existing guest molecules. After pretreatment, the temperature was decreased to 30 C, and N2 was

switched to CO2 for adsorption. For CO2 adsorption, the change in weight over time was recorded for

6 h. Rapid adsorption of CO2 was observed; in particular, the initial adsorption rate was 4.72 wt%/min.

Because of rapid kinetics, approximately 85% of CO2 adsorption equilibrium capacity was attained

within 30 min.

To determine the binding strength between CO2 and PC-2:1-700, the isosteric heat of adsorption (Qst)

was calculated from the CO2 adsorption isotherms at 0 and 25 C by applying the Clausius–Clapeyron

equation [52]. As can be observed from Fig. 11, Qst values were 25–30 kJ/mol, characteristic of typical

physisorption. During physisorption, adsorption and desorption are typically facile, and the adsorbent

is easily regenerated, attributed to weak adsorbent–adsorbate interactions and a low energy requirement

for desorption.

The adsorption–desorption cyclic stability of PC-2:1-700 was further investigated by TGA. PC-2:1-

700 was pretreated at 150 C for 60 min under N2 before the cyclic stability test. During adsorption, the

sample was exposed to CO2 at 30 C for 40 min. During desorption, the sample was purged with N2 at

the same temperature for 60 min to eliminate the adsorbed CO2 molecules. Fig. 12a shows the change

in the weight of PC-2:1-700 during consecutive adsorption and desorption cycles. Rapid kinetics was

observed during adsorption and desorption. The initial adsorption uptake was well maintained during 9

cycles, demonstrating that the adsorbent is easily regenerated by N2 and exhibits excellent adsorption–

desorption cyclic stability. The cyclic stability of PC-2:1-700 was also confirmed in a repeated

adsorption isotherm measurement at 25 C using a volumetric method. During five repeated

measurements, identical isotherms were obtained (Fig. 12b).

13
4. Conclusions

Porous carbon materials with high microporosity were successfully prepared by the activation of

petroleum coke (PC) using KOH as the activating agent. The porosity of PC-based porous carbon

samples was easily controlled by changing the KOH/PC weight ratio and activation temperature. With

increasing KOH/PC ratio and activation temperature, the surface area and total pore volume increased,

reaching up to 2433 m2/g and 1.11 cm3/g for PC-4:1-700, respectively. The maximum CO2 adsorption

uptake was 3.68 mmol/g at 25 C and 1 bar for PC-2:1-700, implying that during activation, an optimal

KOH amount and activation temperature is necessary to prepare the CO2 adsorbent. The adsorbed CO2

amount is not directly related to the surface area or total pore volume, but it is significantly dependent

on the volume of micropores less than 0.8 nm, suggesting that a narrow micropore volume is crucial for

CO2 adsorption. Besides high CO2 uptake, PC-2:1-700 exhibited high selectivity of CO2 over N2, rapid

adsorption–desorption kinetics, and excellent cyclic stability, demonstrating promise as an adsorbent

for CO2 capture.

Acknowledgments

This study was supported by grants from the Basic Science Research Program [grant number

2015R1A1A1A05001363] and the Super Ultra Low Energy and Emission Vehicle Engineering

Research Center [NRF-2016R1A5A1009592] of the National Research Foundation of Korea (NRF)

funded by the Korean Government (Ministry of Science, ICT and Future Planning).

14
References

[1] T.J. Crowley, Causes of climate change over the past 1000 years, Science 289 (2000) 270–277.

[2] R.K. Pachauri, M.R. Allen, V.R. Barros, J. Broome, W. Cramer, R. Christ, J.A. Church, L. Clarke,

Q. Dahe, P. Dasgupta, et al., Climate change 2014: synthesis report, in: R.K. Pachauri, L.A Meyer

(Eds.), Contribution of Working Groups I, II and III to the Fifth Assessment Report of the

Intergovernmental Panel on Climate Change, IPCC, Geneva, Switzerland, 2014, pp. 151.

[3] S.-Y. Lee, S.-J. Park, A review on solid adsorbents for carbon dioxide capture, J. Ind. Eng. Chem.

23 (2015) 1–11.

[4] J.-R. Li, Y. Ma, M.C. McCarthy, J. Sculley, J. Yu, H.-K. Jeong, P.B. Balbuena, H.-C. Zhou, Carbon

dioxide capture-related gas adsorption and separation in metal-organic frameworks, Coordin. Chem.

Rev. 255 (2011) 1791–1823.

[5] S.-M. Hong, K.B. Lee, Solvent-assisted amine modification of graphite oxide for CO2 adsorption,

RSC Adv. 4 (2014) 56707–56712.

[6] A. Wilk, L. Więcław-Solny, A. Tatarczuk, A. Krótki, T. Spietz, T. Chwoła, Solvent selection for

CO2 capture from gases with high carbon dioxide concentration, Korean J. Chem. Eng. 34 (2017) 2275–

2283.

[7] E. Favre, Carbon dioxide recovery from post-combustion processes: can gas permeation membranes

compete with absorption?, J. Membr. Sci. 294 (2007) 50–59.

[8] S.D. Kenarsari, D. Yang, G. Jiang, S. Zhang, J. Wang, A.G. Russell, Q. Wei, M. Fan, Review of

recent advances in carbon dioxide separation and capture, RSC Adv. 3 (2013) 22739–22773.

[9] L. Zhang, N. Xu, X. Li, S. Wang, K. Huang, W.H. Harris, W.K. Chiu, High CO2 permeation flux

enabled by highly interconnected three-dimensional ionic channels in selective CO2 separation

membranes, Energy Environ. Sci. 5 (2012) 8310–8317.

[10] S.S. Hosseini, N. Peng, T.S. Chung, Gas separation membranes developed through integration of

polymer blending and dual-layer hollow fiber spinning process for hydrogen and natural gas

enrichments, J. Membr. Sci., 349 (2010) 156–166.

15
[11] A.R. Millward, O.M. Yaghi, Metal-organic frameworks with exceptionally high capacity for

storage of carbon dioxide at room temperature, J. Am. Chem. Soc. 127 (2005) 17998–17999.

[12] J.A. Mason, K. Sumida, Z.R. Herm, R. Krishna, J.R. Long, Evaluating metal-organic frameworks

for post-combustion carbon dioxide capture via temperature swing adsorption, Energy Environ. Sci. 4

(2011) 3030–3040.

[13] M.R. Hudson, W.L. Queen, J.A. Mason, D.W. Fickel, R.F. Lobo, C.M. Brown, Unconventional,

highly selective CO2 adsorption in zeolite SSZ-13, J. Am. Chem. Soc. 134 (2012) 1970–1973.

[14] B. Guo, L. Chang, K. Xie, Adsorption of carbon dioxide on activated carbon, J. Nat. Gas Chem.

15 (2006) 223–229.

[15] M.S. Shafeeyan, W.M.A.W. Daud, A. Houshmand, A. Arami-Niya, Ammonia modification of

activated carbon to enhance carbon dioxide adsorption: effect of pre-oxidation, Appl. Surf. Sci. 257

(2011) 3936–3942.

[16] N. Hiyoshi, K. Yogo, T. Yashima, Adsorption characteristics of carbon dioxide on organically

functionalized SBA-15, Micropor. Mesopor. Mat. 84 (2005) 357–365.

[17] N.P. Wickramaratne, M. Jaroniec, Activated carbon spheres for CO2 adsorption, ACS Appl. Mater.

Interfaces 5 (2013) 1849–1855.

[18] M. Nandi, K. Okada, A. Dutta, A. Bhaumik, J. Maruyama, D. Derks, H. Uyama, Unprecedented

CO2 uptake over highly porous N-doped activated carbon monoliths prepared by physical activation,

Chem. Commun. 48 (2012) 10283–10285.

[19] A. Ahmadpour, D.D. Do, The preparation of active carbons from coal by chemical and physical

activation, Carbon 34 (1996) 471–479.

[20] W.M.A.W. Daud, W.S.W. Ali, M.Z. Sulaiman, The effects of carbonization temperature on pore

development in palm-shell-based activated carbon, Carbon 38 (2000) 1925–1932.

[21] M.G. Plaza, C. Pevida, C.F. Martín, J. Fermoso, J.J. Pis, F. Rubiera, Developing almond shell-

derived activated carbons as CO2 adsorbents, Sep. Purif. Technol. 71 (2010) 102–106.

16
[22] X. Fan, L. Zhang, G. Zhang, Z. Shu, J. Shi, Chitosan derived nitrogen-doped microporous carbons

for high performance CO2 capture, Carbon 61 (2013) 423–430.

[23] G.K. Parshetti, S. Chowdhury, R. Balasubramanian, Biomass derived low-cost microporous

adsorbents for efficient CO2 capture, Fuel 148 (2015) 246–254.

[24] M.S. Rana, V. Sámano, J. Ancheyta, J.A.I. Diaz, A review of recent advances on process

technologies for upgrading of heavy oils and residua, Fuel 86 (2007) 1216–1231.

[25] B. Jiang, Y. Zhang, J. Zhou, K. Zhang, S. Chen, Effects of chemical modification of petroleum

cokes on the properties of the resulting activated carbon, Fuel 87 (2008) 1844–1848.

[26] M. Yang, L. Guo, G. Hu, X. Hu, J. Chen, S. Shen, W. Dai, M. Fan, Adsorption of CO2 by petroleum

coke nitrogen-doped porous carbons synthesized by combining ammoxidation with KOH activation,

Ind. Eng. Chem. Res. 55 (2016) 757–765.

[27] S. Lowell, J.E. Shields, M.A. Thomas, M. Thommes, Characterization of Porous Solids and

Powders: Surface Area, Pore Size and Density, Kluwer Academic Publishers, Netherlands, 2004.

[28] M.A. Lillo-Ródenas, D. Cazorla-Amorós, A. Linares-Solano, Understanding chemical reactions

between carbons and NaOH and KOH: an insight into the chemical activation mechanism, Carbon 41

(2003) 267–275.

[29] X. Hu, M. Radosz, K.A. Cychosz, M. Thommes, CO2-filling capacity and selectivity of carbon

nanopores: synthesis, texture, and pore-size distribution from quenched-solid density functional theory

(QSDFT), Environ. Sci. Technol. 45 (2011) 7068–7074.

[30] M. Sevilla, C. Falco, M.-M. Titirici, A.B. Fuertes, High-performance CO2 sorbents from algae,

RSC Adv. 2 (2012) 12792–12797.

[31] H. Ding, J.-S. Wei, H.-M. Xiong, Nitrogen and sulfur co-doped carbon dots with strong blue

luminescence, Nanoscale 6 (2014) 13817–13823.

[32] J. Shu, S. Cheng, H. Xia, L. Zhang, J. Peng, C. Li, S. Zhang, Copper loaded on activated carbon

as an efficient adsorbent for removal of methylene blue, RSC Adv. 7 (2017) 14395–14405.

17
[33] D. Sun, R. Ban, P.-H. Zhang, G.-H. Wu, J.-R. Zhang, J.-J. Zhu, Hair fiber as a precursor for

synthesizing of sulfur-and nitrogen-co-doped carbon dots with tunable luminescence properties, Carbon

64 (2013) 424–434.

[34] T. Anjos, A. Charlton, S.J. Coles, A.K. Croft, M.B. Hursthouse, M. Kalaji, P.J. Murphy, S.J.

Roberts-Bleming, Electropolymerization studies on a series of thiophene-substituted 1,3-Dithiole-2-

ones: solid-state preparation of a novel TTF-derivatized polythiophene, Macromolecules 42 (2009)

2505–2515.

[35] J. Heeg, C. Kramer, M. Wolter, S. Michaelis, W. Plieth, W.-J. Fischer, Polythiophene–O3 surface

reactions studied by XPS, Appl. Surf. Sci. 180 (2001) 36–41.

[36] H. Seema, K.C. Kemp, N.H. Le, S.-W. Park, V. Chandra, J.W. Lee, K.S. Kim, Highly selective

CO2 capture by S-doped microporous carbon materials, Carbon 66 (2014) 320–326.

[37] D. Saha, G. Orkoulas, J. Chen, D.K. Hensley, Adsorptive separation of CO2 in sulfur-doped

nanoporous carbons: selectivity and breakthrough simulation, Micropor. Mesopor. Mat. 241 (2017)

226–237.

[38] X. Lu, X. Bian, G. Nie, C. Zhang, C. Wang, Y. Wei, Encapsulating conducting polypyrrole into

electrospun TiO2 nanofibers: a new kind of nanoreactor for in situ loading Pd nanocatalysts towards p-

nitrophenol hydrogenation, J. Mater. Chem. 22 (2012) 12723–12730.

[39] L.K.C. de Souza, N.P. Wickramaratne, A.S. Ello, M.J.F. Costa, C.E.F. da Costa, M. Jaroniec,

Enhancement of CO2 adsorption on phenolic resin-based mesoporous carbons by KOH activation,

Carbon 65 (2013) 334–340.

[40] V. Presser, J. McDonough, S.-H. Yeon, Y. Gogotsi, Effect of pore size on carbon dioxide sorption

by carbide derived carbon, Energy Environ. Sci. 4 (2011) 3059–3066.

[41] H. Wei, S. Deng, B. Hu, Z. Chen, B. Wang, J. Huang, G. Yu, Granular bamboo‐derived activated

carbon for high CO2 adsorption: the dominant role of narrow micropores, ChemSusChem 5 (2012)

2354–2360.

18
[42] M. Sevilla, A.B. Fuertes, Sustainable porous carbons with a superior performance for CO2 capture,

Energy Environ. Sci. 4 (2011) 1765–1771.

[43] H. Cong, M. Zhang, Y. Chen, K. Chen, Y. Hao, Y. Zhao, L. Feng, Highly selective CO2 capture

by nitrogen enriched porous carbons, Carbon 92 (2015) 297–304.

[44] N.P. Wickramaratne, M. Jaroniec, Tailoring microporosity and nitrogen content in carbons for

achieving high uptake of CO2 at ambient conditions, Adsorption 20 (2014) 287–293.

[45] D. Karadag, M. Turan, E. Akgul, S. Tok, A. Faki, Adsorption equilibrium and kinetics of reactive

black 5 and reactive red 239 in aqueous solution onto surfactant-modified zeolite, J. Chem. Eng. Data

52 (2007) 1615–1620.

[46] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.-H. Bae, J.R.

Long, Carbon dioxide capture in metal-organic frameworks, Chem. Rev. 112 (2012) 724–781.

[47] D.M. D'Alessandro, B. Smit, J.R. Long, Carbon dioxide capture: prospects for new materials,

Angew. Chem. Int. Ed. 49 (2010) 6058–6082.

[48] A.L. Myers, J.M. Prausnitz, Thermodynamics of mixed‐gas adsorption, AIChE J. 11 (1965) 121–

127.

[49] Y.-S. Bae, K.L. Mulfort, H. Frost, P. Ryan, S. Punnathanam, L.J. Broadbelt, J.T. Hupp, R.Q. Snurr,

Separation of CO2 from CH4 using mixed-ligand metal-organic frameworks, Langmuir 24 (2008) 8592–

8598.

[50] D.W. Hand, S. Loper, M. Ari, J.C. Crittenden, Prediction of multicomponent adsorption equilibria

using ideal adsorbed solution theory, Environ. Sci. Technol. 19 (1985) 1037–1043.

[51] D. Lee, C. Zhang, C. Wei, B.L. Ashfeld, H. Gao, Hierarchically porous materials via assembly of

nitrogen-rich polymer nanoparticles for efficient and selective CO2 capture, J. Mater. Chem. A 1 (2013)

14862–14867.

[52] S. Chowdhury, R. Mishra, P. Saha, P. Kushwaha, Adsorption thermodynamics, kinetics and

isosteric heat of adsorption of malachite green onto chemically modified rice husk, Desalination 265

(2011) 159–168.

19
Table 1

Elemental analysis of vacuum residue (VR), petroleum coke (PC), and PC-derived porous carbon

samples.

Elemental composition (wt%)


Sample
C H N O S

VR 79.75 9.37 0.47 1.48 4.20

PC 83.53 3.34 1.02 1.25 7.47

PC-1:1-700 83.84 0.89 0.63 10.93 0.85

PC-2:1-700 80.70 1.54 0.32 13.58 1.19

PC-3:1-700 75.17 1.97 0.24 19.91 2.86

PC-4:1-700 85.04 1.00 0.21 7.83 0.35

PC-2:1-600 72.30 1.40 1.01 14.85 2.42

PC-2:1-800 86.66 0.32 0.25 5.46 0.53

20
Table 2

Textural properties of the PC-derived porous carbon samples.

Sample SBET a (m2/g) Vt b (cm3/g) V0 c (cm3/g) VN d (cm3/g)

PC-1:1-700 915 0.38 0.37 0.22

PC-2:1-700 1433 0.60 0.58 0.31

PC-3:1-700 1586 0.66 0.64 0.30

PC-4:1-700 2433 1.11 0.86 0.28

PC-2:1-600 538 0.22 0.21 0.16

PC-2:1-800 1600 0.70 0.62 0.26


a
Specific surface area calculated according to the BET method from the N2 adsorption data at relative
pressures between 0.01 and 0.1. bTotal pore volume obtained from the amount of N2 adsorbed at a
relative pressure of 0.99. cMicropore volume determined by the Dubinin–Radushkevich equation.
d
Narrow micropore volume for pore size below 0.8 nm calculated by the non-local density functional
theory model.

21
Table 3

CO2 adsorption uptake of the PC-derived porous carbon samples at different temperatures and at 1 bar.

CO2 adsorption uptake (mmol/g)


Sample
0 C 25 C 50 C

PC-1:1-700 4.56 2.92 1.73

PC-2:1-700 6.08 3.68 2.14

PC-3:1-700 6.08 3.57 1.99

PC-4:1-700 6.12 3.43 1.83

PC-2:1-600 3.80 2.62 1.70

PC-2:1-800 4.89 3.17 1.79

22
Table 4

Parameters from the Langmuir model fitting for the adsorption of CO2 on PC-2:1-700.

H (= qmax  b,
Temperature (C) qmax (mmol/g) b (1/bar) R2
mmol/g∙bar)

0 9.97 1.493 14.89 0.9972

25 7.47 0.943 7.05 0.9987

50 5.21 0.681 3.54 0.9994

23
Figure captions

Fig. 1. SEM images of (a) pristine PC, (b) PC-1:1-700, (c) PC-2:1-700, (d) PC-3:1-700, (e) PC-4:1-700, (f)

PC-2:1-600, and (g) PC-2:1-800. (h) HRTEM image of PC-2:1-700.

Fig. 2. N2 adsorption (solid symbols) and desorption (open symbols) isotherms at 77 K for PC-derived

porous carbon samples (a) with different KOH/PC ratios and (b) with different activation temperatures.

Fig. 3. (a) Pore size distribution and (b) cumulative pore volume of PC-derived porous carbon samples with

different KOH ratios. (c) Pore size distribution and (d) cumulative pore volume of PC-derived porous

carbon samples with different activation temperatures.

Fig. 4. Thermal decomposition of PC mixed with KOH (the ratio of KOH and PC was 2:1).

Fig. 5. FT-IR spectra of pristine PC and PC-derived porous carbon samples.

Fig. 6. XPS spectra of pristine PC and PC-derived porous carbon samples: (a) S2p and (b) N1s.

Fig. 7. CO2 adsorption isotherms at 25 C for PC-derived porous carbon samples (a) with different KOH

ratios and (b) with different activation temperatures. Relationship between CO2 uptake at 25 C and 1 bar

and textural properties: (c) BET specific surface area and (d) volume of micropores less than 0.8 nm.

Fig. 8. CO2 adsorption isotherms of PC-2:1-700 at 0, 25, and 50 C. Dots and solid lines represent

experimental data and Langmuir model fitting data, respectively.

Fig. 9. (a) CO2 and N2 adsorption isotherms of PC-2:1-700 at 25 C. (b) The IAST selectivity of PC-2:1-

700 for CO2 over N2 at CO2/N2 = 15:85 at 25 C.

Fig. 10. CO2 adsorption kinetics of PC-2:1-700 at 30 C and 1 bar.

Fig. 11. Isosteric heat of adsorption for PC-2:1-700 calculated from CO2 adsorption isotherms.

Fig. 12. Cyclic stability test of PC-2:1-700 from (a) a gravimetric method at 30 C and (b) a volumetric

method at 25 C.

24
Fig. 1. SEM images of (a) pristine PC, (b) PC-1:1-700, (c) PC-2:1-700, (d) PC-3:1-700, (e) PC-4:1-700, (f)

PC-2:1-600, and (g) PC-2:1-800. (h) HRTEM image of PC-2:1-700.

25
Fig. 2. N2 adsorption (solid symbols) and desorption (open symbols) isotherms at 77 K for PC-derived

porous carbon samples (a) with different KOH/PC ratios and (b) with different activation temperatures.

26
Fig. 3. (a) Pore size distribution and (b) cumulative pore volume of PC-derived porous carbon samples with

different KOH ratios. (c) Pore size distribution and (d) cumulative pore volume of PC-derived porous

carbon samples with different activation temperatures.

27
Fig. 4. Thermal decomposition of PC mixed with KOH (the ratio of KOH and PC was 2:1).

28
Fig. 5. FT-IR spectra of pristine PC and PC-derived porous carbon samples.

29
Fig. 6. XPS spectra of pristine PC and PC-derived porous carbon samples: (a) S2p and (b) N1s.

30
Fig. 7. CO2 adsorption isotherms at 25 C for PC-derived porous carbon samples (a) with different KOH

ratios and (b) with different activation temperatures. Relationship between CO2 uptake at 25 C and 1 bar

and textural properties: (c) BET specific surface area and (d) volume of micropores less than 0.8 nm.

31
Fig. 8. CO2 adsorption isotherms of PC-2:1-700 at 0, 25, and 50 C. Dots and solid lines represent

experimental data and Langmuir model fitting data, respectively.

32
Fig. 9. (a) CO2 and N2 adsorption isotherms of PC-2:1-700 at 25 C. (b) The IAST selectivity of PC-2:1-

700 for CO2 over N2 at CO2/N2 = 15:85 at 25 C.

33
Fig. 10. CO2 adsorption kinetics of PC-2:1-700 at 30 C and 1 bar.

34
Fig. 11. Isosteric heat of adsorption for PC-2:1-700 calculated from CO2 adsorption isotherms.

35
Fig. 12. Cyclic stability test of PC-2:1-700 from (a) a gravimetric method at 30 C and (b) a volumetric

method at 25 C.

36

You might also like