You are on page 1of 33

Journal Pre-proofs

Full Length Article

Natural Honeycomb-like Structure Cork Carbon with Hierarchical Micro-


Mesopores and N-containing Functional Groups for VOCs Adsorption

Xiang Xu, Yang Guo, Rui Shi, Hongyu Chen, Yankun Du, Baogen Liu,
Zheng Zeng, Ziyu Yin, Liqing Li

PII: S0169-4332(21)01620-2
DOI: https://doi.org/10.1016/j.apsusc.2021.150550
Reference: APSUSC 150550

To appear in: Applied Surface Science

Received Date: 9 March 2021


Revised Date: 8 June 2021
Accepted Date: 3 July 2021

Please cite this article as: X. Xu, Y. Guo, R. Shi, H. Chen, Y. Du, B. Liu, Z. Zeng, Z. Yin, L. Li, Natural
Honeycomb-like Structure Cork Carbon with Hierarchical Micro-Mesopores and N-containing Functional
Groups for VOCs Adsorption, Applied Surface Science (2021), doi: https://doi.org/10.1016/j.apsusc.2021.150550

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. 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.

© 2021 Published by Elsevier B.V.


Natural Honeycomb-like Structure Cork Carbon with

Hierarchical Micro-Mesopores and N-containing Functional

Groups for VOCs Adsorption


Xiang Xu a, Yang Guo a, Rui Shi a, Hongyu Chen a, Yankun Du a, Baogen Liu a, Zheng

Zeng a *, Ziyu Yin b *, Liqing Li a *

a School of Energy Science and Engineering, Central South University, Changsha,

410083, China

b Department of Nanoscience, Joint School of Nanoscience and Nanoengineering

(JSNN), University of North Carolina at Greensboro, Greensboro, NC 27401, United

States

E-mail address:

zengzheng@csu.edu.cn,

z_yin@uncg.edu,

liqingli@hotmail.com

1 / 32
Abstract

It can be said that most of the garbage were misplaced resources. And how to turn these

resources into valuable products is extremely important. In this study, the cork wine

stoppers were recovered and prepared into VOCs adsorption materials through

ammonia activation. The prepared cork biomass carbon had abundant micropores and

mesopores and rich N-containing functional groups, especially amino groups (3.6%).

The obtained sample NC900 exhibited a good performance for the adsorption of

acetone (1221 mg g-1 at 18 kPa), benzene (840 mg g-1 at 10 kPa) and toluene (720 mg

g-1 at 3 kPa) at 25 °C. The adsorption effect of hierarchical porous carbon on VOCs

under different pressures were calculated and analyzed through GCMC. Multilayer

adsorption promoted the adsorption of VOCs molecules in mesopores under the high

pressure. The promotion effects of nitrogen-containing functional groups (amino,

pyrrolic, pyridine) based on adsorption energy calculations were evaluated via DFT.

The amino functional groups obtained the strongest affinity for acetone (-0.64 eV),

benzene (-0.66 eV) and toluene (-0.72 eV). This study achieved a guiding significance

for the preparation of cork biomass carbon and its application in the VOCs adsorption.

Keywords: Cork; Porous carbon; VOCs; Adsorption; GCMC; DFT

2 / 32
Introduction

Before Robert Hooker first observed cork cells, cork has already attracted attention over

centuries for its softness and breathability.[1] It was usually used to make wine stoppers,

crafts, heat insulation boards and sound insulation boards. This is probably due to the

excellent elasticity, sealing, thermal insulation, sound insulation, electrical insulation

and friction resistance given by its unique honeycomb-like structure.[2] About 201

kilotons of cork were annually produced from the worldwide cork oak forests.[3]

Meanwhile, a large amount of consumed cork products (such as wine corks) were sent

to the garbage station each year, which means approaches to utilize these waste corks

were urgently needed. Recently, these waste corks gradually attracted attention.[4]

Thanks to their abundant natural pore structure, corks can be considered for adsorbing

pollutants as well.[5, 6] However, current reports mainly discussed their potential

application in adsorbing heavy metals and drugs in sewage, while rare reports focus on

the gas adsorption. Therefore, it is of great significance to explore the conversion of

waste corks and promote the application in air pollutant adsorption.

Volatile organic compounds (VOCs), a class of air pollutants, cause the recognized

hazard to atmospheric environment and human health.[7] The treatments of VOCs were

generally based on recovery (such as absorption, adsorption, membrane separation, and

condensation) and destruction (such as thermal, catalytic, and biological oxidation)

methods.[8, 9] It is worth mentioning that the performance of removing VOCs will be

better when other technologies are combined with the adsorption technology. In view

3 / 32
of the advantages of this method, scholars have carried out a lot of researches on the

removal of VOCs by the adsorption method.[10] And the adsorption materials include

activated carbon, CNT, zeolite, silica gel, clay, metal organic framework (MOF),

hypercrosslinked polymer (HCP), macroporous polymer, composite materials and other

porous materials.[11-21] In order to save costs, researchers prefer economical materials

with abundant raw materials and low prices.[22] Therefore, the use of waste biomass

resources seems to provide us an option.

VOCs adsorption depends on the aperture structure, specific surface area and surface

functional groups. And the biomass carbon with high specific surface area, abundant

hierarchical pores and rich functional groups were pursued for a better performance.[11,

23] Herein, activated carbon from waste cork stoppers was prepared by ammonia gas

treatment, and its adsorption performance for acetone, benzene and toluene was tested.

Physical and chemical characterization were performed to investigate the properties of

amination activated cork carbon. Furthermore, the VOCs adsorption mechanism on the

carbon surface was analyzed by computational chemistry. In detail, micropores and

mesopores were investigated through Grand canonical Monte Carlo (GCMC) and the

contribution of N-containing functional groups was analyzed via density functional

theory (DFT) calculations.

Experimental

Synthesis

4 / 32
These corks in this study were collected from Liangyuanziniang winery. The cork

stoppers were pulverized into small particles (< 900 μm, 20 mesh sieving) for 5 min via

a high-speed multifunctional crusher (800Y, Kemanshi), and pre-carbonized in N2 (100

mL min-1) at 800 °C for 1 hour with 5 °C min-1 programmed heating rate. And the pre-

carbonized cork was washed with 30 mL hydrochloric acid (1.0 M), and then washed

with distilled water until the pH was neutral. The washed carbon was dried in a vacuum

oven at 60 °C for 12 hours and named as C. And the per-carbonized material C was

conducted by the programmed temperature (5 °C min-1) in the furnace under N2 flow

(100 mL min-1) and activated for 1 hour under NH3 (100 mL min-1) at 700 °C, 800 °C

and 900 °C, and named as NC700, NC800 and NC900, respectively. The yield of pre-

carbonized carbon C from waste cork was approximately 16.3%, and the yield of

NC700, NC800 and NC900 were 11.3%, 9.9% and 5.6%, respectively.

Characterization

N2 and VOCs adsorption isotherms were performed on a vapor adsorption analyzer

instrument (JW-BK132Z, Beijing JWGB Sci & Tech Co., Ltd). N2 adsorption was

carried out at -196.15 °C, while acetone, benzene and toluene adsorption were tested at

25 °C. The pore size distribution and specific surface area of cork carbon were

calculated on the non-local density functional theory (NLDFT) and Brunauer-Emmett-

Teller (BET) method from N2 adsorption isotherms.[24] And the Vmicro (< 2 nm) and

Vmes+mar (2 nm -300 nm) were analyzed via the cumulative results from NLDFT method.

Microstructure images of samples were collected by a scanning electron microscope

5 / 32
(SEM) (JSM-7900F, JEOL) under 20 kV acceleration voltage. And micromorphology

of samples was observed on a transmission electron microscopy (TEM) (Titan G2 60-

300, FEI). N-containing functional groups were investigated through an X-ray

photoelectron spectroscopy (XPS) (ESCALAB 250Xi, ThermoFisher). The collected

XPS data was referenced to the adventitious carbon peak (C 1s 284.8 eV).

Adsorption isotherm models

Freundlich and dual-site Langmuir-Freundlich (DSLF) models were used to describe

the adsorption isotherms,[25, 26] Freundlich model was shown in Equation 1:

qF = KFp1/nF (1)

where KF means Freundlich affinity constant (mmol g−1 atm−1/n), p means the total bulk

pressure (kPa) and nF means the heterogeneity factor that represents a deviation from

the linearity of adsorption.

The DSLF model was shown in Equation 2:


1/n1 1/n2
b1p b2p
qDSLF = q1 1/n1 + q2 1/n2 (2)
1 + b1p 1 + b2p

where q was the model parameter of saturation capacity at sites (1 or 2) (mmol g−1),

and b and n are the affinity coefficients and the deviations from the ideal homogeneous

surface, respectively.[27, 28]

Calculations

GCMC simulations were carried out to investigate the VOCs adsorption process for

various pore sizes. The carbon model used in this calculation was simplified into a slit

pore model with parallel single layer graphite sheets. All the GCMC simulations

6 / 32
contained 107 steps to ensure the equilibration at each state point, and followed by

2×107 steps to determine the average production. The Lennard-Jones (LJ) parameters

of VOCs were taken from the COMPASS force field.[29] The adsorption capacities

with different pressures were obtained through a series of simulations at a specific

temperature of 25 °C.

DFT calculations were carried out via the Perdew-Burke-Ernzerhof (PBE) of the

generalized gradient approximation (GGA).[30] The tolerances of energy, gradient, and

displacement convergence were 1 × 10−5 Ha, 2 × 10−3 Ha/Å, and 5 × 10−3 Å,

respectively. And the vacuum slab was set as 16 nm to minimize the interaction between

the nearest slabs. Therefore, the VOCs adsorption energies were calculated by the

Equation 3:

Ead = EVOCs@sur - (Esur + EVOCs) (3)

where EVOCs@sur represents the total energy of adsorbent and the VOCs molecule on its

surface, and Esur and EVOCs represent the energy of adsorbent and free VOCs molecule,

respectively.[31]

Results and Discussion

Cork stoppers were pulverized (the brown powder in Scheme 1) and pre-carbonized at

800 °C. The elemental analysis and ICP-MS were performed to investigate the

elemental contents of the cork stoppers (Table S1 and Table S2). The acid pickling was

conducted to avoid the irons influence from adsorption in this research. In order to gain

7 / 32
the amino carbon and developed pores, the prepared carbon C was amination activated

from 700 °C to 900 °C. As shown in the SEM images, all the samples revealed the

honeycomb-like structure, which resulted from the nature structure of cork cell (Figure

1 and S1). And the unique honeycomb structure of cork was severely etched after

ammonia activation at 900 °C (Figure 1d). This means that NH3 etching has contributed

to the development of macroporous structures and the increase of interconnected cells,

which will facilitate the mass transfer. However, it is necessary to further evaluate the

microporous and mesoporous structure of these materials to investigate their effects on

the VOCs adsorption.

Scheme 1. Synthesis of ammoniated carbon from waste cork stoppers.

8 / 32
Figure 1. SEM images of C (a), NC700 (b), NC800 (C), and NC900 (d).

Figure 2. N2 adsorption (a) and pore size distribution (b) of samples.

To investigate the pore development in NH3 amination, N2 adsorption was performed

to analyze the micropore and mesopore of samples. The adsorption and desorption

curves of C, NC700 and NC800 exhibited I type adsorption according to IUPAC

classification, while NC900 showed Ⅰ type combined with IV type isotherms. This

9 / 32
indicated that sample NC900 has hierarchical pores with various mirco-mesopores

while other samples were ascribed to micropores. And pore size distribution (PSD) was

obtained from NLDFT model (Figure 2b). Micropore volume was increased in the

amination process and further increased in a higher temperature (from 700 °C to

900 °C). In addition, mesopores were significantly developed in NC900. This thanks to

the effects of thermal instability and NH3 cavitation. To analyze the contribution of

NH3 cavitation, sample C was activated in N2 at 900 °C as a comparison (named as

C900). N2 adsorption-desorption isotherm of C900 was shown in Figure S2. The PSD

figure demonstrated that both thermal instability and amination could promote the

formation of micropore and mesopore, however the mesopore (especially 2-4 nm) was

mainly developed via amination (Figure S3). The pore size development of the sample

before and after ammonia gas activation at 700 °C, 800 °C and 900 °C were observed

by high-resolution TEM (Figure 3). The development of micropores and mesopores

proved the NH3 activation process at a higher temperature, especially at 900 °C. The

BET specific surface area was calculated and reported in Table 1. The pre-carbonized

C showed an uncompetitive specific surface area of 369 m2 g-1 and small micropore

volume of 0.17 cm3 g-1. And the specific surface area of the sample has a significant

increase after ammonia activation, in detail the specific surface area increases from 558

m2 g-1 to 2060 m2 g-1 as the activation temperature increases from 700 °C to 900 °C.

Therefore, NH3 activation of cork carbon was benefit for the development of

hierarchical pores and the increase of specific surface area.

10 / 32
Table 1. Physical properties of samples.

Method Name BET (m2 g-1) Vmicro (cm3 g-1) Vmes+mar (cm3 g-1)

Pre-carbonized C 369 0.17 0.06

N2 C900 1149 0.52 0.44

NC700 558 0.26 0.10

NH3 NC800 1022 0.46 0.22

NC900 2060 0.75 1.46

Figure 3. High-resolution TEM images of C (a), C900 (b), NC700 (c), NC800 (d) and

NC900 (e).

11 / 32
Figure 4. XPS analysis of N functional groups. (a) C, (b) NC700, (c) NC800, (d) NC900,

and (e) N species and atomic ratio.

In addition to the analysis of the pore structure, the N-containing functional groups on

the carbon surface after ammonia activation should be surveyed as well. XPS was

performed to investigate the N species after amination. The deconvolution of N 1s

spectra were shown in Figure 4a-d. Sample C without NH3 etching showed oxidized-N

(402.0 eV), graphitic-N (401.0 eV), pyrrolic-N (399.8 eV) and pyridinic-N (398.6 eV).

And amino (400.5 eV) were detected in NC700, NC800 and NC900 after NH3 treating,

while oxidized-N and graphitic-N were disappeared after amination. The atomic ratio

showed that amino was increased along the treated temperature (Figure 4e). In addition,

the content of pyridinic-N in aminated samples was higher than that of pyrrolic-N,

which reaches agreement with the higher thermal stability of pyridinic-N.[32] Note that,

the pyrrolic-N and pyridinic-N were slightly reduced in NC800 compared to NC700,
12 / 32
which was due to the instability in a high temperature (over 800 °C). Followed by the

increasing of amination temperature (900 °C), the transform of pyrrolic-N and amino

into pyridinic-N led to the pyridinic-N enhancement (Figure 5a).[33, 34] Totally, N

species were increased after amination and especially the atom fraction of nitrogen has

reached 6.44% in NC900. And the increasing of N content resulted from the NH3

activation process, which caused the transform of other functional groups such as O-

containing functional groups and the higher consumption of carbonaceous material

around 900 °C.[35] Therefore, the ammonia activation could be benefit for the

development of hierarchical micro-mesopores and the production of amino groups

(Figure 5b). And N-containing functional groups were observed to be evenly distributed

on the carbon surface in the elemental maps (Figure 5c).

Figure 5. Development of hierarchical pores and transform of N-containing functional

groups in the NH3 amination process.

13 / 32
Table 2. Physical parameters of acetone, benzene and toluene.

Vapor Dipole Molar


Kinetic
Adsorbate Formula pressure moment ×1018 volume Structure Ref.
diameter(Å)
(kPa) (esu cm) (cm3 mol-1)

Acetone C3H6O 4.60 32.03 2.880 209.0 [36]

Benzene C6H6 5.35-5.85 11.74 0.000 256.0 [36]

Toluene C7H8 5.25 3.28 0.375 316.0 [36]

In order to explore the application of aminated cork carbon in the gas adsorption,

acetone, benzene and toluene, three of the most common and toxic non-halogenated

compounds, were chosen as the VOCs adsorbate.[37] These target VOCs were the most

common VOCs in industrial air pollutant. In addition, recent measurement and

monitoring reports have indicated that the concentration of benzene and toluene

increased noticeably after the natural fires.[38] The common information of these

VOCs was collected in Table 2, which demonstrated the different polarities and

dynamic diameters of acetone, benzene and toluene.[36] The VOCs adsorption

isotherms of samples at 25 °C were reported in Figure 6a-c. The maximum VOC

adsorption capacity of acetone (ca. 1221 mg g-1 at 18 kPa), benzene (ca. 840 mg g-1 at

10 kPa) and toluene (ca. 720 mg g-1 at 3 kPa) were obtained by NC900, which was far

larger than that of other samples (Figure S4). Note that, the adsorption isotherms of C,
14 / 32
NC700 and NC800 have reached the nearly saturated adsorption capacity below 1 kPa,

while that of NC900 reveled a long rang raising with the increasing of pressure. This

was usually related to the pore size of the materials, introduced below. The Table S3

ensured the great VOCs adsorption performance of NC900 compared with other

researches.[22, 39-43]

Figure 6. VOCs adsorption isotherms (a-c) and Freundlich/DSLF fitting (d-f).

Figure 7. Recovery tests of NC900 for acetone (a), benzene (b) and toluene (c).
15 / 32
Experimental data was fitted by Freundlich and DSLF model (Figure 6d-f), and the

fitting data was reported in Table 3. The multilayer adsorption was implied due to the

good fitting of Freundlich model. From the R2 values, DSLF model had a good fitting

on the adsorption isotherms of acetone, benzene and toluene. Therefore, both of

Freundlich and DSLF models can describe and predict the adsorption behavior of this

three VOCs well. And the increasing of n parameters implies that the sample NC900

has a heterogeneous surface after ammonia activation. The adsorption selectivity for

acetone, benzene and toluene vapors were calculated via the ideal adsorbed solution

theory (IAST) method (Table S4). Moreover, the low-pressure acetone, benzene and

toluene adsorption isotherms of C, NC700, NC800 and NC900 were fitted through the

Langmuir model (Figure S5).

Furthermore, in order to evaluated the commercial application potential of amination

activated cork, the acetone cycling adsorption was conducted to investigated the

recovery ability. The regeneration and resorption process were test in the N2 flow at

105 °C and 7000 ppm acetone, benzene and toluene flow at 25 °C, respectively. After

five regeneration cycles, the resorption ability was decayed less than 10% (Figure 7).

This demonstrated NC900 performed an excellent regeneration ability towards the

adsorption of these three kinds of VOCs.

16 / 32
Table 3. Fitting parameters of Freundlich and DSLF models for VOCs adsorption

isotherms.

Freundlich DSLF
Sorbent Sorbate
KF nF R2 q1 b1 n1 q2 b2 n2 R2

Acetone 44.78 3.59 0.871 63.94 5.00 0.45 128.59 0.01 0.77 0.928

C Benzene 61.85 5.07 0.891 40.72 0.01 0.49 83.18 5.00 0.76 0.988

Toluene 49.78 3.44 0.926 64.68 5.00 0.88 162.44 0.01 0.11 0.975

Acetone 59.63 3.84 0.935 105.76 0.01 0.54 87.97 5.00 0.70 0.984

NC700 Benzene 102.89 5.44 0.891 79.66 2.74 1.30 79.66 2.73 1.30 0.926

Toluene 75.34 3.31 0.859 54.14 5.00 0.67 54.14 5.00 0.67 0.949

Acetone 86.33 4.09 0.930 127.58 5.00 0.81 80.17 0.01 0.54 0.985

NC800 Benzene 118.81 5.02 0.902 237.21 0.01 0.33 166.73 5.00 0.86 0.980

Toluene 120.34 3.63 0.765 85.04 5.00 0.80 85.04 5.00 0.80 0.869

Acetone 237.03 1.77 0.999 1500.00 0.08 1.36 1500.00 0.08 1.36 0.994

NC900 Benzene 430.41 3.30 0.956 272.70 0.86 0.50 1500.00 0.24 2.46 0.994

Toluene 515.60 2.95 0.876 380.45 4.25 0.68 380.45 4.25 0.68 0.988

Pore size was usually considered to be one of the main factors to determine the

adsorption ability. The aminated NC900, which has revealed brilliant VOCs adsorption

capacities, exhibited unique mesopores. Therefore, except micropore, the VOCs

adsorption at mesopores should be discussed as well. In this study, GCMC was

17 / 32
performed to investigate the VOCs adsorption among carbon sheet with different pore

size (0.8 nm-5.0 nm) at 25 °C. Due to the adsorption limitation of the pore size, the pore

size below the size of the adsorbate was excluded. Herein, the pore size of calculated

carbon sheet was started from 0.8 nm.

Figure 8. VOCs adsorption capacities at fixed pressure (a-c) and acetone adsorption

density (red points) in pore size from 0.8 nm to 5.0 nm (d).

18 / 32
Acetone, benzene and toluene adsorption capacity were calculated at 18 kPa, 10 kPa

and 3 kPa, respectively (Figure 8 a-c). The adsorption capacity of acetone, benzene and

toluene reached 11.57 mmol mL-1, 10.24 mmol mL-1 and 8.45 mmol mL-1, respectively.

Note that, the adsorption capacity of the three VOCs revealed a trend of first increasing,

then stabilizing and followed by rapidly decreasing as the increase of pore sizes. For

instance, the adsorption capacity of acetone increases with the expansion of the pore

size from 0.8 nm to 1.6 nm. From 1.6 nm to 3.2 nm, the saturated adsorption capacity

revealed slightly changes, and the saturated adsorption capacity decreases rapidly with

the further expansion of the pore size from 3.2 nm. The size range from 1.6 nm to 3.2

nm could be considered as more effective pores for acetone adsorption at 18 kPa, and

that from 1.6 nm to 4.5 nm for benzene and toluene at corresponding pressure.

To better understood the interesting changes of VOCs adsorption with pore sizes, the

densities of acetone adsorption were analyzed and displayed at 2 kPa, 10 kPa and 18

kPa (Figure 8d). The adsorption density at 0.8 nm and 1.2 nm indicated a monolayer

adsorption for each carbon sheet, and the multilayer adsorption occurred from 1.6 nm.

Since the pore size was enlarged to a turning point (ca. 2.0 nm), the ratio of actual

adsorption pore volume to total pore volume gradually decreases at 2 kPa (Figure 8d).

And this was also observed at 10 kPa and 18 kPa. Moreover, the turning point seems to

move to a larger pore size as the increasing of pressure. This means that when the

pressure increases from a lower level, micropores and even mesopores will play a role

in the VOCs adsorption. Especially, the turning point at 18 kPa was obtained at between

19 / 32
3.2 nm and 3.6 nm, which is consistent with the point where the adsorption capacity

drops (Figure 8a). This demonstrated that the acetone effective adsorption pore volume

ratio will gradually decrease from 3.6 nm. This gives a reasonable explanation why the

VOCs adsorption capacity gradually decreases with the further expansion of the pore

size.

In addition, as the pressure increases from 2 kPa to 10 kPa and 18 kPa, the distance

between carbon sheet and acetone molecule increases from 0.60 nm to 0.89 nm and

1.04 nm, respectively (Figure 8d). Due to the interaction between adsorbate and

adsorbent, acetone molecules preferentially adsorb on the surface of the carbon sheet,

forming a monolayer adsorption. As the pressure increases, acetone molecules have

gradually accumulated on the surface of the carbon sheet, and some acetone molecules

even began to adsorb on the surface of the previous layer of acetone molecules, that is,

multilayer adsorption. The distance did not change as the expansion of pore size at a

fixed pressure. At this point, the turning point of the adsorption pore size corresponding

to the pressure can be further understood. Since the increase in pore size provided

favorable conditions for multilayer adsorption, acetone molecules could interact with

the carbon surface and acetone molecules as well. Furthermore, when the acetone

molecules adsorbed by the upper carbon and have a close interaction distance with that

adsorbed by the lower carbon sheet, it is beneficial to adsorb extra acetone molecules

layers in the middle. Therefore, all the adsorbed acetone molecular layers contributed

to the aperture tuning point of the adsorption capacity. However, the extra acetone

20 / 32
molecules layers would disappear when the pore size was further increased. Due to the

reduction of the acetone adsorption layer, an ineffective adsorption space appeared

between the slit pores, resulting in a rapid decline in the adsorption capacity of acetone

after the turning point.

Figure 9. Calculated adsorption isotherm of acetone at 2.8 nm pore.

Acetone adsorption capacity was calculated at 2.8 nm (Figure 9), which was the most

observed mesopore in the pore size distribution of NC900. The snapshot image of

acetones at 6 kPa indicated a monolayer adsorption. And the acetone adsorption images

at 9 kPa demonstrated the presence of a second layer, which means the beginning of

multilayer adsorption. Therefore, the adsorption of acetone at 2.8 nm slit pores was

ascribed to the monolayer adsorption before 6 kPa and multilayer adsorption starting

from around 10 kPa. This could be the reason that experimental acetone adsorption

isotherm of NC900 remains the increasing trending at around 10 kPa (Figure 6a).

21 / 32
Figure 10. The most stable adsorption configurations of VOCs (a) acetone, (b) benzene,

(c) toluene on the surface of amino, pyrrolic, pyridine and none functional groups

carbon.

Except pore size effect, the functional group on the carbon surface was regarding as

one of the most important issues for VOCs adsorption. In this research, N-containing

functional groups were evaluated based on DFT calculations. All the N-containing

functional groups showed a higher adsorption energy for VOCs than that on pure carbon

model, and the highest adsorption energy of acetone (-0.64 eV), benzene (-0.66 eV) and

toluene (-0.72 eV) demonstrated amino revealed the best enhancement for VOCs

adsorption. Thanks to the hydrogen bond interaction, the affinity of amino groups and

acetone molecules is stronger than that of pyridine and pyrrole (Figure 10a). Due to the

22 / 32
polarity of the acetone molecule, the stronger polar functional group (amino) exhibits a

much better affinity than other N-containing functional groups. Benzene and toluene

were generally considered to be adsorbed on the carbon surface by π-π interaction

between carbon sheet and aromatic rings (Figure 10b and c).[44] In addition, we found

that some of the nitrogen-containing functional groups changed the structure of the

carbon sheet, resulting in the ups and downs of the carbon sheet, and affecting the

surface electronic structure of the carbon sheet. Therefore, the N-containing functional

groups in the carbon sheet resulted in a stronger promotion of adsorption energy than

that on the edge of the carbon sheet in our previous report.[43, 45]

Conclusion

This research provided a new utilization strategy for waste cork (wine stoppers). The

high specific surface area (2060 m2 g-1) biomass carbon NC900 with rich micro-

mesopores was synthesized by ammonia activation at 900 °C. NC900 had the highest

amino functional group content (3.6%) compared with other samples. NC900 exhibited

the best VOCs adsorption performance 1221 mg g-1 for acetone at 18 kPa, 840 mg g-1

for benzene at 10 kPa and 720 mg g-1 for toluene at 3 kPa. GCMC calculations

demonstrated that multilayer adsorption promoted the adsorption of VOCs molecules

in mesopores under the high pressure. Turning point of the most effective pore size

range for acetone adsorption expanded from 2.0 nm to 3.2 nm as the increasing of

pressure from 2 kPa to 18 kPa. The DFT results proved that N-containing functional

23 / 32
groups promoted the adsorption of VOCs on the carbon surface, and the amino

functional groups had the strongest affinity for acetone, benzene and toluene.

Acknowledgements

This research was supported by the National Natural Science Foundation of China (No.

21878338), National Key Research and Development Program of China (Nos.

2019YFC0214302, 2019YFC0214303) and the Key Research and Development

Project of Hunan Province (No.2018SK2038).

Declaration of Interests

The authors declare no competing interests.

24 / 32
References

[1] S.P. Silva, M.A. Sabino, E.M. Fernandes, V.M. Correlo, L.F. Boesel, R.L. Reis,
Cork: properties, capabilities and applications, Int. Mater. Rev., 50 (2005) 345-365.
[2] L. Gil, Cork Composites: A Review, Materials, 2 (2009).
[3] I.M. Aroso, A.R. Araújo, R.A. Pires, R.L. Reis, Cork: Current Technological
Developments and Future Perspectives for this Natural, Renewable, and Sustainable
Material, ACS Sustainable Chem. Eng., 5 (2017) 11130-11146.
[4] Y. Li, Y. Lu, Q. Meng, A.C.S. Jensen, Q. Zhang, Q. Zhang, Y. Tong, Y. Qi, L. Gu,
M.-M. Titirici, Y.-S. Hu, Regulating Pore Structure of Hierarchical Porous Waste Cork-
Derived Hard Carbon Anode for Enhanced Na Storage Performance, Adv. Energy
Mater., 9 (2019) 1902852.
[5] Q. Wang, Z. Lai, J. Mu, D. Chu, X. Zang, Converting industrial waste cork to
biochar as Cu (II) adsorbent via slow pyrolysis, Waste Manag., 105 (2020) 102-109.
[6] J.A.C. Castellar, J. Formosa, A.I. Fernández, P. Jové, M.G. Bosch, J. Morató, H.
Brix, C.A. Arias, Cork as a sustainable carbon source for nature-based solutions treating
hydroponic wastewaters – Preliminary batch studies, Sci. Total Environ., 650 (2019)
267-276.
[7] M. Kampa, E. Castanas, Human health effects of air pollution, Environ. Pollut., 151
(2008) 362-367.
[8] M.S. Kamal, S.A. Razzak, M.M. Hossain, Catalytic oxidation of volatile organic
compounds (VOCs) – A review, Atmos. Environ., 140 (2016) 117-134.
[9] W. Zou, B. Gao, Y.S. Ok, L. Dong, Integrated adsorption and photocatalytic
degradation of volatile organic compounds (VOCs) using carbon-based
nanocomposites: A critical review, Chemosphere, 218 (2019) 845-859.
[10] L. Zhu, D. Shen, K.H. Luo, A critical review on VOCs adsorption by different
porous materials: Species, mechanisms and modification methods, J. Hazard. Mater.,
389 (2020) 122102.
[11] X. Ma, R. Chen, K. Zhou, Q. Wu, H. Li, Z. Zeng, L. Li, Activated Porous Carbon
with an Ultrahigh Surface Area Derived from Waste Biomass for Acetone Adsorption,
CO2 Capture, and Light Hydrocarbon Separation, ACS Sustainable Chem. Eng., 8
(2020) 11721-11728.
[12] B. Liu, R. Shi, R. Chen, C. Wang, K. Zhou, Y. Ren, Z. Zeng, Y. Liu, L. Li,
Optimized synthesis of nitrogen-doped carbon with extremely high surface area for
adsorption and supercapacitor, Appl. Surf. Sci., 538 (2021) 147961.
[13] W. Shi, D.L. Plata, Vertically aligned carbon nanotubes: production and
applications for environmental sustainability, Green Chem., 20 (2018) 5245-5260.
[14] S. Jafari, F. Ghorbani-Shahna, A. Bahrami, H. Kazemian, Adsorptive removal of
toluene and carbon tetrachloride from gas phase using Zeolitic Imidazolate Framework-
8: Effects of synthesis method, particle size, and pretreatment of the adsorbent,
Microporous Mesoporous Mater., 268 (2018) 58-68.
[15] H. Sui, H. Liu, P. An, L. He, X. Li, S. Cong, Application of silica gel in removing

25 / 32
high concentrations toluene vapor by adsorption and desorption process, J. Taiwan Inst.
Chem. Eng., 74 (2017) 218-224.
[16] Y. Wang, X. Su, Z. Xu, K. Wen, P. Zhang, J. Zhu, H. He, Preparation of surface-
functionalized porous clay heterostructures via carbonization of soft-template and their
adsorption performance for toluene, Appl. Surf. Sci., 363 (2016) 113-121.
[17] K. Yang, Q. Sun, F. Xue, D. Lin, Adsorption of volatile organic compounds by
metal–organic frameworks MIL-101: Influence of molecular size and shape, J. Hazard.
Mater., 195 (2011) 124-131.
[18] J. Wang, W.-Q. Wang, Z. Hao, G. Wang, Y. Li, J.-G. Chen, M. Li, J. Cheng, Z.-
T. Liu, A superhydrophobic hyper-cross-linked polymer synthesized at room
temperature used as an efficient adsorbent for volatile organic compounds, RSC Adv.,
6 (2016) 97048-97054.
[19] T. Zhu, Q. Yu, L. Ding, T. Di, T. Zhao, T. Li, L. Li, Atom-economical preparation
of polybismaleimide-based microporous organic polymers, Green Chem., 21 (2019)
2326-2333.
[20] L. Jia, W. Yu, C. Long, A. Li, Adsorption equilibrium and dynamics of gasoline
vapors onto polymeric adsorbents, Environ. Sci. Pollut. Res., 21 (2014) 3756-3763.
[21] Y. Zheng, F. Chu, B. Zhang, J. Yan, Y. Chen, Ultrahigh adsorption capacities of
carbon tetrachloride on MIL-101 and MIL-101/graphene oxide composites,
Microporous Mesoporous Mater., 263 (2018) 71-76.
[22] Y. Du, H. Chen, X. Xu, C. Wang, F. Zhou, Z. Zeng, W. Zhang, L. Li, Surface
modification of biomass derived toluene adsorbent: hierarchically porous
characterization and heteroatom doped effect, Microporous Mesoporous Mater., 293
(2020) 109831.
[23] X. Ma, L. Li, Z. Zeng, R. Chen, C. Wang, K. Zhou, C. Su, H. Li, Synthesis of
nitrogen-rich nanoporous carbon materials with C3N-type from ZIF-8 for methanol
adsorption, Chem. Eng. J., 363 (2019) 49-56.
[24] J.W.F. To, J. He, J. Mei, R. Haghpanah, Z. Chen, T. Kurosawa, S. Chen, W.-G.
Bae, L. Pan, J.B.H. Tok, J. Wilcox, Z. Bao, Hierarchical N-Doped Carbon as CO2
Adsorbent with High CO2 Selectivity from Rationally Designed Polypyrrole Precursor,
J. Am. Chem. Soc., 138 (2016) 1001-1009.
[25] Y. Chao, B. Tang, J. Luo, P. Wu, D. Tao, H. Chang, X. Chu, Y. Huang, H. Li, W.
Zhu, Hierarchical porous boron nitride with boron vacancies for improved adsorption
performance to antibiotics, J. Colloid Interface Sci., 584 (2021) 154-163.
[26] Z. Xiang, X. Peng, X. Cheng, X. Li, D. Cao, CNT@Cu3(BTC)2 and Metal–Organic
Frameworks for Separation of CO2/CH4 Mixture, J. Phys. Chem. C, 115 (2011) 19864-
19871.
[27] D. Li, L. Li, R. Chen, C. Wang, H. Li, H. Li, A MIL-101 Composite Doped with
Porous Carbon from Tobacco Stem for Enhanced Acetone Uptake at Normal
Temperature, Ind. Eng. Chem. Res., 57 (2018) 6226-6235.
[28] X. Peng, Q. Jin, Ideal adsorbed solution theory, two-dimensional equation of state,
and molecular simulation for separation of H2/N2/O2/CH4/CO in graphite nanofiber and

26 / 32
C60 intercalated graphite, Sep. Purif. Technol., 237 (2020) 116369.
[29] Y. An, Q. Fu, D. Zhang, Y. Wang, Z. Tang, Performance evaluation of activated
carbon with different pore sizes and functional groups for VOC adsorption by
molecular simulation, Chemosphere, 227 (2019) 9-16.
[30] J.P. Perdew, J.A. Chevary, S.H. Vosko, K.A. Jackson, M.R. Pederson, D.J. Singh,
C. Fiolhais, Atoms, molecules, solids, and surfaces: Applications of the generalized
gradient approximation for exchange and correlation, Phys. Rev. B, 46 (1992) 6671-
6687.
[31] K. Zhou, W. Ma, Z. Zeng, X. Ma, X. Xu, Y. Guo, H. Li, L. Li, Experimental and
DFT study on the adsorption of VOCs on activated carbon/metal oxides composites,
Chem. Eng. J., 372 (2019) 1122-1133.
[32] K. Stańczyk, R. Dziembaj, Z. Piwowarska, S. Witkowski, Transformation of
nitrogen structures in carbonization of model compounds determined by XPS, Carbon,
33 (1995) 1383-1392.
[33] J.R. Pels, F. Kapteijn, J.A. Moulijn, Q. Zhu, K.M. Thomas, Evolution of nitrogen
functionalities in carbonaceous materials during pyrolysis, Carbon, 33 (1995) 1641-
1653.
[34] X. Zhou, L. Qiu, R. Fan, A. Wang, H. Ye, C. Tian, S. Hao, Y. Yang, Metal–
Organic Framework-Derived N-Rich Porous Carbon as an Auxiliary Additive of Hole
Transport Layers for Highly Efficient and Long-Term Stable Perovskite Solar Cells,
Sol. RRL, 4 (2020) 1900380.
[35] Y. Ji, S. Wang, Y. Dong, L. Li, L. Li, X. Feng, X. Lu, L. Mu, Tuning nitrogen
species on natural biomass derived porous carbon for efficient acetone adsorption,
Mater. Chem. Phys., 253 (2020) 123338.
[36] J.-R. Li, R.J. Kuppler, H.-C. Zhou, Selective gas adsorption and separation in
metal–organic frameworks, Chem. Soc. Rev., 38 (2009) 1477-1504.
[37] L.F. Liotta, Catalytic oxidation of volatile organic compounds on supported noble
metals, Appl. Catal., B, 100 (2010) 403-412.
[38] L.A. Simms, E. Borras, B.S. Chew, B. Matsui, M.M. McCartney, S.K. Robinson,
N. Kenyon, C.E. Davis, Environmental sampling of volatile organic compounds during
the 2018 Camp Fire in Northern California, J. Environ. Sci., 103 (2021) 135-147.
[39] K. Zhou, W. Ma, Z. Zeng, R. chen, X. Xu, B. Liu, H. Li, H. Li, L. Li, Waste
biomass-derived oxygen and nitrogen co-doped porous carbon/MgO composites as
superior acetone adsorbent: Experimental and DFT study on the adsorption behavior,
Chem. Eng. J., 387 (2020) 124173.
[40] R. Chen, N. Han, L. Li, S. Wang, X. Ma, C. Wang, H. Li, H. Li, L. Zeng,
Fundamental understanding of oxygen content in activated carbon on acetone
adsorption desorption, Appl. Surf. Sci., 508 (2020) 145211.
[41] B. Rubahamya, K.S. Kumar Reddy, A. Prabhu, A. Al Shoaibi, C. Srinivasakannan,
Porous carbon screening for benzene sorption, Environ. Prog. Sustainable Energy, 38
(2019) S93-S99.
[42] H.-L. Chiang, K.-H. Lin, C.-Y. Chen, C.-G. Choa, C.-S. Hwu, N. Lai, Adsorption

27 / 32
Characteristics of Benzene on Biosolid Adsorbent and Commercial Activated Carbons,
J. Air Waste Manage. Assoc., 56 (2006) 591-600.
[43] C. Su, Y. Guo, H. Chen, J. Zou, Z. Zeng, L. Li, VOCs adsorption of resin-based
activated carbon and bamboo char: Porous characterization and nitrogen-doped effect,
Colloids Surf., A, 601 (2020) 124983.
[44] H. Wang, J. Gao, X. Xu, B. Liu, L. Yu, Y. Ren, R. Shi, Z. Zeng, L. Li, Adsorption
of Volatile Organic Compounds (VOCs) on Oxygen-rich Porous Carbon Materials
Obtained from Glucose/Potassium Oxalate, Chem. - Asian J., 16 (2021) 1118-1129.
[45] C. Su, Y. Guo, L. Yu, J. Zou, Z. Zeng, L. Li, Insight into specific surface area,
microporosity and N, P co-doping of porous carbon materials in the acetone adsorption,
Mater. Chem. Phys., 258 (2021) 123930.

28 / 32
Highlights

 High specific surface area porous carbon was obtained from waste cork.

 NC900 had the highest N-containing functional group content and amino groups.

 NC900 exhibited the best adsorption performance for acetone, benzene and toluene.

 Multilayer adsorption promoted the adsorption in mesopores via GCMC

calculations.

 Amino had the strongest affinity for VOCs through DFT calculations.

29 / 32
Graphical abstract

30 / 32
Declaration of interests

☒ The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in this
paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:

31 / 32
Credit Author Statement

X.X., L.L. and Z.Z. conceptualized the experiments and synthesized the samples. X.X.,

Y.G. and H.C. performed the first-principles calculations. X.X., R.S., Y.D. and B.L.

collected and analyzed the characterization and adsorption data. L.L., Z.Y. and Z.Z.

supervised the project. X.X. wrote the manuscript. All authors proofread, commented

on, and approved the final manuscript for submission.

32 / 32

You might also like