You are on page 1of 25

Journal Pre-proofs

Full Length Article

Separation of azeotrope acetone-methanol mixture at high-concentration con-


ditions: Experimental and theoretical perspective of the molecular packing ef-
fects in mesopores

Yang Guo, Changqing Su, Hongyu Chen, Jinxian Wang, Baogen Liu, Zheng
Zeng, Liqing Li

PII: S0169-4332(23)00790-0
DOI: https://doi.org/10.1016/j.apsusc.2023.157113
Reference: APSUSC 157113

To appear in: Applied Surface Science

Received Date: 8 February 2023


Revised Date: 20 March 2023
Accepted Date: 21 March 2023

Please cite this article as: Y. Guo, C. Su, H. Chen, J. Wang, B. Liu, Z. Zeng, L. Li, Separation of azeotrope
acetone-methanol mixture at high-concentration conditions: Experimental and theoretical perspective of the
molecular packing effects in mesopores, Applied Surface Science (2023), doi: https://doi.org/10.1016/j.apsusc.
2023.157113

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.

© 2023 Published by Elsevier B.V.


Separation of azeotrope acetone-methanol mixture at high-concentration
conditions: Experimental and theoretical perspective of the molecular
packing effects in mesopores
Yang Guoa, Changqing Sub, Hongyu Chena, Jinxian Wanga, Baogen Liua , Zheng Zenga,* and Liqing Lia,*

aSchool of Energy Science and Engineering, Central South University, Changsha 410083, Hunan, China

bSchool of Resources and Environment, Carbon Neutralization Research Insitute, Hunan University of
Technology and Business, Changsha 410205, Hunan, China

* Corresponding author.

E-mail address: hdzengzheng@163.com (Zheng Zeng)

E-mail address: liqingli@hotmail.com (Liqing Li)

Abstract

The recycling and concentration of acetone-methanol mixture organic solvents are very
important to the chemical industry and biologically related processes. Hence, the separation
properties of microporous and mesoporous activated carbons (MIC and MEC) for the acetone-
methanol mixture were investigated for the first time by combining experiments and
I
theoretical calculations. The results showed that the calculated dynamic acetone (methanol)
adsorption capacities of MIC and MEC are 1.60 (10.66) and 5.15 (6.01) mmol/g at high
pressures, respectively. MEC exhibited a greater methanol/acetone selectivity (6.66) than MIC
(1.17). Subsequent theoretical calculations studied the adsorption behaviors of acetone and
methanol molecules from the perspective of the adsorbate-framework as well as adsorbate-
adsorbate interactions and found the different molecular structure of acetone and methanol
was the main factor in their different separation preferences. Notably, the molecular packing
effects associated with three molecular packing steps in the separation were further clarified
by the radical distribution functions and self-created slit model. The 2.0-3.0 nm mesopore had
more unfilled pore spaces after adsorbing a layer of acetone molecules, and thus had a
significant selectivity towards methanol molecules.

Keywords: Activated carbons; Pore sizes; Methanol/acetone; Separation; GCMC&DFT.

1. Introduction

Acetone, one of the most widely polar solvents in the chemical industry, is generally used
as a common component of mixed organic solvents for its improved solvation properties (e.g.,
dielectric constant, density, etc.)[1],[2],[3]. Acetone-methanol mixture is the simplest example
of such solvent mixtures, which easily diffuse into the atmosphere and cause harm to humans
and the environment[4],[5],[6]. Typically, acetone cannot be easily separated from methanol
by distillation for the formation of the azeotropic system[7]. The development of effective
strategies in the separation and capture of the acetone-methanol mixture is of great significance
in the recycling and concentration of solvent mixtures.

The physical adsorption based on weak interactions has vital significance in the field of
gas capture and separation for its spontaneous and reversible process[8],[9],[10]. Numerous
porous materials are developed as adsorbents for gas capture and separation, among them,
activated carbons (ACs) have been considered of great industrial interest due to their eco-
friendly property and low cost[11],[12]. The separation performance of the ACs is directly
related to their surface chemistry properties. The formation of surface heterogeneity can
provide additional active adsorption sites by turning the electronic properties of the interaction
surface [13],[14]. For example, the introduced oxygen, nitrogen and sulfur-containing
functional groups adsorb methanol preferentially due to strong gas-framework electrostatic
interactions, resulting in a high methanol/acetone selectivity of AC materials[15],[16],[17].
An important point to note is that surface modification can only influence the interaction
between adsorbent surface and adsorbate, and thus, mainly affects the adsorption and
2
separation performance of ACs at low pressure (mainly monolayer adsorption).

For the separation of organic gas mixtures with low air volume and high concentration in
industrial practice, the pores of the AC adsorbents are nearly or completely saturated[18].
Hence, the pore structure of AC material is a key factor affecting the adsorption and separation
of gas mixture at high pressures. For example, compared to mesoporous carbons, microporous
carbons are more effective in the adsorption and separation of light hydrocarbons[19]. For now,
however, rare researchers have systematically investigated the pore size distribution effects of
AC materials on their acetone and methanol adsorptive separation performance. Besides, the
separation of gas mixtures in AC materials under pore saturation conditions is mainly dictated
by molecular packing effects, which favor the selective of molecules with higher packing
efficiencies [20],[21]. Therefore, the packing efficiencies of acetone and methanol, dominated
by the gas-framework interaction as well as the gas-gas interactions, are needed to be clarified.

To address the above issues, we combined adsorption experiments and theoretical


calculations (Grand canonical Monte Carlo (GCMC), Molecular Dynamics (MD) and density
functional theory (DFT) to reveal the influence of pore size on the acetone-methanol mixture
separation performance of activated carbons. Note that methanol and acetone, as common
organic solvents, are often mixed in different molar ratios for practical applications. In this
study, we take the molar ratio of 50:50 of methanol and acetone in the binary organic solvent
mixture as an example, and assume that the molar ratio of methanol and acetone in the vapor
mixture evaporated from this binary mixture system remains constant. In terms of
experimental and simulation techniques, GCMC simulations and dynamic breakthrough
experiments were carried out to investigate the separation performance of the microporous and
mesoporous carbons. Then the adsorption behaviors of acetone and methanol were clarified
by adsorbate-framework and adsorbate-adsorbate interactions, weak interaction analysis and
radial distribution functions (RDFs). Lastly, the molecular packing effects and mechanisms
underlying the co-adsorption process of the acetone-methanol mixture were clarified by the
visualized results of the GCMC simulations. This study is of great scientific interest in both
theory and practice, which also provides a novel insight for designing ACs for the effective
acetone-methanol mixture adsorptive separation.

2. Experiment and simulation

2.1. Chemicals and materials

Acetone (C3H6O, >99%), methanol (CH3OH, >99,8%) and potassium hydroxide

3
(KOH, >99.9%) were obtained from Sinopharm Reagent Co., Ltd. The microporous and
mesoporous activated carbons (MIC and MEC) derived from coconut shell (commercial
activated carbon purchased from Jiangxi Province) were washed with ultrapure water to
remove any fine particles, followed by drying at 120 °C for 18 hours in an oven.

2.2. Characterization

The activated carbons were subjected to nitrogen adsorption and desorption isotherm
measurements using the JW-ZQ200C instrument (Beijing JWGB Technology Co., Ltd., China)
at 77 K. The specific surface area of the material was determined using the multi-point BET
method. Prior to testing, both MIC and MEC were vacuum pre-treated at 200°C for 4 hours.
The pore size distribution (PSD) of the porous carbon was obtained using the non-local density
functional theory (NLDFT) method based on the nitrogen adsorption and desorption isotherms
measured at 77 K. The micromorphology of the sample was observed using a transmission
electron microscope (TEM, Titan G2 60-300, FEI). The crystal phases of MIC and MEC were
characterized using a Raman spectrometer (LabRam Hr800) and X-ray diffractometer (XRD,
PANalytical, The Netherlands). The surface elemental composition was determined using X-
ray photoelectron spectroscopy (XPS, Thermo Fisher Scientific Inc., USA).

2.3. Adsorption experiment

The dynamic breakthrough experiments of acetone-methanol were performed using a


self-made fixed-bed adsorption reactor, which was equipped with a mass spectrometer (MS,
OmniStar GSD 320, Pfeiffer Vacuum) at the outlet to monitor the VOCs concentration. Before
the test, 100 mg of activated carbon sample was weighed and placed in the reactor. The reactor
was then heated to 100°C and purged with nitrogen for 3 hours. The dynamic adsorption
amounts of VOCs of MIC and MEC were calculated using the following Eq.1 [22]:

F t
qt = m ∫0(C0 - C) dt (1)

Here, qt represents the uptake amounts of VOCs (mmol/g) on the samples, F refers to the flow
rate (25 mL/min) of the mixed gas, m is the weight of MIC or MEC. C0 is the initial
concentration (ppm, volume fraction) of the VOCs mixture gas at the system inlet, and C is
the outlet concentration.

The acetone/methanol selectivity (S) of MEC and MIC was evaluated by ideal adsorption
solution theory (IAST), shown in Eq.2[23]:
4
q1 p2
S = q2 × p1 (2)

Where q1 and q2 represent the mole fractions of methanol and acetone in the adsorbed
phase, while p1 and p2are the corresponding mole fractions in the bulk phase.

2.4 Model and methods

To study the effect of micropores and mesopores of the AC materials, the slit pore model
formed by parallel graphite sheets was used for the molecular simulations. Note that the
simplified nature of the slit model may result in an incomplete depiction of the intricate porous
structure in actual activated carbon, potentially leading to a simplification of the interplay
between the adsorbate and the porous surface. However, considering the difficulty of modeling
the disordered structure of amorphous activated carbons, the slit model remains a powerful
tool for predicting the adsorption behavior of activated carbons and gaining insight into the
underlying mechanisms of adsorption. Detailed information about the model can be found in
the Supplementary Material (Fig. S1 and Eq. S1). GCMC calculations were carried out to
evaluate the adsorption and separation performance of single or mixed gas components in the
slit pore models. All the GCMC calculations were performed in the simulation software
RASPA [24],[25]. The GCMC simulations were began with 2.0×106 cycles of equilibration
steps, followed by 4.0×106 cycles of production steps. The van der Waals (vdW) and
coulombic interactions were calculated by LJ (12-6) [26] model and Ewald summation method,
respectively. The potential energy of the gas-sorbate as well as the gas-gas system was
calculated by the sum of the electrostatic and the vdW energy, which was defined as Eq.3:

σij 12 σij 6
U(rij) = 4εij [( ) - ( ) ] +
rij rij
qiqj
rij
(3)

Where 𝑟𝑖𝑗 denotes the distance between atoms i and j, εij and σij denote energy and size
parameters, respectively. All the interaction parameters conform to Lorentz–Berthelot mixing
rules, εij = (εii ∙ εjj)1/2 , σij = (σii + σjj)/2. 𝑞𝑖 is the quantity of electric charges of atom 𝑖.

The LJ potential parameters for the C in the slit pore models were taken from the Dreiding
force field [27], while the acetone and methanol molecules were described by the transferable
potentials for phase equilibria (TraPPE-UA) force field[28],[29]. The LJ parameters and point
charges for the gas molecules were presented in Table S1. In the TraPPE-UA force field, the
methyl groups in acetone and methanol are considered to be united-atom located on the carbon
5
atom site, which has been widely used in predicting the separation behavior in AC [15],[30]
and MOFs[31].

3. Result

In this section, the GCMC simulations were initially carried out to estimate the optimal
separation pore size for the acetone and methanol mixture. Then the dynamic breakthrough
experiments were carried out to experimentally study the methanol/acetone selectivity of the
microporous and mesoporous activated carbon.

3.1. The optimal separation pore size calculated by GCMC simulations

The co-adsorption of acetone-methanol mixture (mole bulk composition 50:50) in the slit
pore models with the pore size of 0.8-4.0 nm at 293 K was established by GCMC simulations.
Fig.1 displays the selectivity of acetone and methanol mixture in the slit pore models at 2 kPa
and 20 kPa, respectively.

Fig. 1. GCMC simulations for equimolar acetone/methanol mixtures in the slit pore models
at (a) 2 kPa and (b) 20 kPa.

As shown in Fig. 1a, all the activated carbon pore size (0.8-4.0 nm) exhibit acetone
selectivity at 2 kPa. The selectivity (acetone/methanol) decreases as pore size increases, and
the selected micropores (pore size < 2.0 nm) have a better acetone/methanol selectivity
(ranging from 5.7 to 4.0) at 2 kPa. However, as the acetone-methanol mixture pressure
increases to 20 kPa (Fig. 1b), the activated carbon shows an obvious methanol adsorption
preference in the 0.8-4.0 nm pore size range. The mesopores (pore size 2.0-4.0 nm) exhibit a
6
higher methanol/acetone selectivity (~22.1-13.2). Especially, the mesopores being 3.0 nm
show the largest methanol uptake amount (19.3 mmol/cm3) and selectivity (~22.1).

3.2. Characterization of microporous and mesoporous activated carbon samples

According to the calculation results in Fig. 1, we can confirm that the micropores and
mesopores had a different effect on the separation performance for the acetone-methanol
mixture at low and high pressures. To further explore the difference between microporous and
mesoporous activated carbon in the separation performance of acetone-methanol mixture, two
porous carbons (named MIC and MEC) with different pore characteristics were selected to
study their adsorption performance for acetone and methanol.

The porous structure of MIC and MEC was characterized by N2 adsorption/desorption


isotherms at 77 K (Fig. 2a). MIC shows typical I-type curves, suggesting that the samples were
typical microporous materials [32]. However, the isotherms of MEC exhibit the combined
characteristics of type Ⅰ and type IV isotherms [33],[34], and the obvious hysteresis loop
presents a H4 character (0.4 ≤ P/P0 < 0.99), showing the existence of abundant mesopores with
continuous pore size distribution. The detailed pore size distribution (PSD) information
calculated by using the NLDFT method for MIC and MEC is displayed in Fig. 2b. MIC shows
concentrated PSD in 0.8-2.0 nm, while MEC presents a relatively wide pore size distribution
in the mesopores range (2.0-4.0 nm). Fig. 2c shows the pore volume distribution parameters.
Their micropore volume is respectively 0.62 and 0.32 cm3/g, with the corresponding mesopore
volumes of 0.19 and 0.60 cm3/g. In addition, the pore size of MIC and MEC were observed by
high-resolution TEM (Fig. S3). The TEM results demonstrate the abundant microporous pore
channels on the surface of MIC. On the other hand, more mesopores can be observed in MEC
compared to MIC, which is consistent with the results of PSD.

7
Fig. 2. Characterization of MIC and MEC. (a) Nitrogen adsorption−desorption isotherms. (b)
Pore size distribution (c) Pore volume for N2 adsorption. (d) XPS survey. (e) XRD patterns.
(f) Raman spectra.

In addition, the surface chemical properties of the two samples were measured by XPS
(Fig. 2d). The XPS survey spectra of MIC and MEC exhibit two main peaks at around 284.8
and 532.9 eV, corresponding to the C1s and O1s core level, respectively (Fig. 2d) [35],[36].
MIC and MEC exhibit a similar oxygen content (5.7 vs. 5.1 at%). The detailed XPS spectra
for O1 s of MIC and MEC had been examined as well (see Fig. S2). To further analyze the
structure of the MIC and MEC, the XRD measurement was implemented. As shown in Fig.
2e, the two broad peaks at about 2θ values of 25° and 44° are attributed to the (002) and (100)
crystal facets of graphite, respectively[37]. In the absence of any other sharp peaks in the range
of 20-80°, the samples are considered to be mainly amorphous. The Raman spectrum of MIC
and MEC displayed in Fig. 2f further corroborates the XRD results. Two characteristic peaks
are observed at 1350 cm-1 and 1590 cm-1, corresponding to D-band (crystal defects) and G-
band (hexagonal graphite), respectively. The ID/IG value for MIC and MEC is about (1.01 vs.
1.03), respectively, indicating a low degree of graphitization.

Overall, MIC and MEC exhibit significant differences in porous structure (micropores
and mesopores dominated, respectively), but have similar BET surface area and elemental
8
characteristics, which minimize the influence of other factors. From the above point of view,
MIC and MEC are competent for the research needs of this study.

3.3. Acetone-methanol separation performance of the MIC and MEC

The dynamic breakthrough experiments were carried out at 293 K to establish the
separation performance of MIC and MEC. Different inlet concentrations of acetone-methanol
mixture were set to investigate the separation preference of the activated carbons at both low
and high pressures. Note that ppm in the text refers to volume fractions. The breakthrough
curve for the empty bed is shown in the Supplementary Material (Fig. S4).

Fig. 3. The breakthrough curves of acetone-methanol mixture in MIC and MEC at (a-b) 8085
ppm and (c-d) 101060 ppm.
9
As displayed in Fig.3a-b, when the inlet concentration of acetone and methanol are both
8065 ppm, both MIC and MEC exhibit acetone selectivity. The dynamic acetone (methanol)
uptake amounts of MIC and MEC are 3.10 (1.72) and 2.40 (1.64) mmol/g, respectively.
Considering the same partial pressure of methanol and ethanol, MIC exhibits a higher
acetone/methanol selectivity than MEC (1.80 vs. 1.46) in the co-adsorption experiments,
which suggests that the microporous activated carbon has a better acetone selectivity than
mesoporous activated carbon at low pressures. Fig. 3c-d shows the breakthrough curves of
MIC and MEC at high-concentration conditions (101060 ppm). The amount of methanol
captured from acetone-methanol mixture in MIC and MEC is 6.01 and 10.66 mmol/g,
respectively, which compares favorably with their acetone uptake amount (5.15 and 1.60
mmol/g). For high-concentration conditions (101060 ppm), MEC exhibits a greater
methanol/acetone selectivity (6.66) than MIC (1.17). The above result suggests mesoporous
carbons are more efficient for acetone-methanol mixture separation at high pressures.

4. Discussion

The above simulation and experimental results demonstrate that microporous and
mesoporous carbon exhibit distinctive acetone-methanol selectivity performance at low and
high pressures. To understand the mechanisms underlying the co-adsorption process, further
simulations were conducted.

4.1. Study of the adsorption behavior of acetone and methanol on microporous and
mesoporous carbons

The co-adsorption of acetone and methanol mixtures on porous carbons is a complicated


problem. Instead, by analyzing the single-component adsorption behavior of acetone and
methanol, we can have a better understanding of the separation mechanism. Fig. S5 displays
the simulated adsorption isotherms of the acetone and methanol on activated carbon with
various pore sizes (0.8-4.0 nm) at 293 K. Obviously, the micropores have a significantly higher
acetone uptake capacity in the entire pressure range while mesopores (2.0-3.0 nm) exhibit a
greater saturated adsorption capacity for methanol. Additionally, static adsorption experiments
were also carried out to investigate the adsorption performance of acetone and methanol vapor
on MIC and MEC (Fig. S6). MIC exhibits a higher acetone adsorption performance under the
entire pressure range, while MEC shows a higher methanol adsorption capacity at high
pressures. The experimental results qualitatively agree with the theoretical simulations,
proving the reliability of our results.

10
Fig. 4. Potential energies of (a-b) acetone and (d-f) methanol on 0.8 and 3.0 nm slit pores in
four cases: the gas-framework van der Waals interactions (VIs), the gas-framework
(electrostatic interactions) EIs, the gas-gas VIs and the gas-gas EIs.

To understand the adsorption behavior of acetone and methanol, the potential energies of
adsorbing the gas molecules on 0.8 and 3.0 nm split pores were divided into four interaction
cases: (i) the gas-framework van der Waals interactions (VIs), (ii) the gas-framework
electrostatic interactions (EIs), (iii) the gas-gas VIs, and (iv) the gas-gas EIs. Note that the
framework refers to the pore walls of the slit pore models. For example, the gas-framework
VIs and gas-framework EIs refer to van der Waals and electrostatic interactions between
acetone or methanol and the pore walls, respectively. As shown in Fig. 4, the potential energy
11
between the framework and the two adsorbates are both dominated by VIs at low pressures.
Specifically, the VI energies between the acetone and frameworks reach a high level (-53.7
and -27.5 kJ/mol in 0.8 and 3.0 nm pores, respectively) at 0.02 kPa, which are remarkably
higher than that of methanol (-28.1 and -15.3 kJ/mol in 0.8 and 3.0 nm pores respectively).
The above results indicate that the interaction between the pore wall and acetone is stronger
than that between the pore wall and methanol, leading to the acetone selectivity of the activated
carbon material at low pressure.

Fig. 5 The isosurfaces mapping of weak interaction for acetone and methanol in (a-b) 0.8 nm
and (c-d) 3.0 nm silt pore.

To provide a more intuitive understanding of the interaction between the gas molecules
and frameworks at low pressures, the interaction analysis was calculated by the Multiwfn
package [38],[39]. The optimal adsorption configurations were established by density
functional with dispersion correction (DFT-D3) in the CP2K package [40]. The detailed
simulation methods and more perspectives for the isosurfaces mapping are presented in the
12
Supplementary Material (Fig. S7 and S8). The isosurfaces in Fig. 5 depict where weak
interactions occur. The red, blue, and green colors represent the strong repulsive, strong
electrostatic interactions, and van der Waals interaction, respectively.

As displayed in Fig. 5, green isosurfaces are observed in all the interfacial regions,
exhibiting the VIs between the gas molecules (acetone and methanol) and the pore walls.
Herein, we speculate that the larger VIs energies between the acetone and pore walls should
be attributed to different structures of methanol and acetone. Specifically, as shown in Fig. 5c
and 5d, the two methyl groups and one carbonyl group of acetone molecule interact strongly
with the pore wall, which shows a significant larger van der Waals interaction isosurfaces than
methanol (with only one methyl group). While in the 0.8 nm silt pores (Fig. 5a and 5b), the
acetone and methanol molecules are attracted by both pore walls. This provides an intuitive
understanding of why the interaction energies between adsorbate molecules and the pore walls
are larger for micropores than that for mesopores. Moreover, due to the structural differences
between acetone and methanol, the micropores exhibit a stronger adsorption promotion effect on
acetone than methanol. The above results together explain why micropores have stronger acetone
selectivity than mesopores.

Fig. 6. The RDFs for O_ace – CH3_ace and O_meth – H_meth distances for (a) acetone and (b)
methanol.

The potential energy of gas-gas interactions (both acetone and methanol) increases with the
increase of pressures (Fig. 4c, 4d). The intermolecular interaction of methanol is stronger than
that of acetone. The EIs between methanol molecules even dominate the adsorption process in
13
3.0 nm slit pore at high pressures. The stronger intermolecular interactions between methanol
molecules lead to their higher packing efficiencies at higher pressures and hence favor the
selective adsorption of methanol molecules. To further confirm the specific interaction
between the molecules, the radical distribution functions (RDFs) for distances between all
atoms of the acetone and methanol molecules in 3.0 nm were established. The RDFs describe
the relative average concentration of neighbors at a distance r from a given molecule. Detailed
RDFs methodology used in this study is presented in Supplementary Material (Fig. S9). Note
that for ease of description, O_ace and CH3_ ace denote the O atom and the methyl group of
acetone, respectively. And O_meth and H_meth represent the methanol molecule's O atom
and H atom, respectively. The RDFs distances for the O_ace - CH3_ ace and the O_meth -
H_meth pairs exhibit the minimum distance in the acetone and methanol molecule,
respectively (see Fig. S10). By combining the gas-gas interactions in Fig. 4 and the RDFs
results, we can determine the packing modes for the two gas molecules at high pressures.
Specifically, As shown in Fig. 6, the RDFs for distances between O and H atoms exhibit a first
peak at an intermolecular distance of 0.18 nm, which is the characteristic of hydrogen bonding
(a cyan isosurface between the hydroxyl group). As for acetone, the RDFs of O and CH3
distances for acetone molecules show a peak at 0.35 nm. The intermolecular interaction
between acetone molecules is mainly contributed by VI and EI between methyl and carbonyl
groups. Note that although there is a C-HO hydrogen bond between the methyl and carbonyl
group of acetone molecules, it is too weak (show the green isosurfaces) and may only be
attributed to van der Waals interaction. Moreover, the self-diffusivities of methanol in 3.0 nm
silt are lower than that of acetone (see Fig. S11), which also proves that the methanol molecules
are more compactly arranged and exhibit a higher packing efficiency than acetone at high
pressures[18],[41].

4.2. Study of the co-adsorption behavior of acetone and methanol on microporous and
mesoporous carbons

The co-adsorption behavior was first studied by visualization approaches. The density
distribution contour of the equimolar acetone-methanol mixture on the slit pores model at 2
and 20 kPa was established by the GCMC simulations. From dark blue, and light blue to green
adsorption sites, the adsorption density increases in sequence [42]. Here, we constructed a new
slit model (named 0.8-3.0 nm slit model) that included both micropores (0.8,1.2 and 1.6 nm)
and mesopores (2.0, 2.5 and 3.0 nm). This allows a more intuitive interpretation of the
adsorption preferences of the micro and mesopores for acetone and methanol molecules at one
certain adsorption condition. Detailed information about the 0.8-3.0 nm split model is
14
described in the Supplementary Material (Fig. S12).

15
Fig. 7. Density distribution contours of (a) acetone and (b) methanol on silt pores model in the
co-adsorption process at 20 kPa. (c) The RDFs for distances O_meth - H_meth and O_ace -
H_meth distances for acetone-methanol mixture in 0.8-3.0 nm silt pore models at 20 kPa. (d)
16
Stable adsorption configurations and Eads of the two adsorbate systems: acetone-acetone and
acetone-methanol.

As shown in Fig. S13, acetone molecules occupied most of the adsorption sites on the
pore walls (especially in the 0.8 to 1.2 nm micropore) by their stronger vdW interaction at 2
kPa. The obvious competitive effects lead to acetone selectivity in both the micro and
mesopores of the activated carbon at low pressures. With the increase of pressure (from 2 to
20 kPa), as shown in Fig.7a, b, most acetone molecules are mainly adsorbed on the pore walls,
showing the feature of monolayer adsorption, while methanol molecules appear multilayer
adsorption and gradually fill the unoccupied pore space. Snapshots of the acetone-methanol
mixture in the silt pore model at 20 kPa provide a more intuitive presentation of this
phenomenon (see Fig. S14). The acetone molecule adsorbed on the pore wall becomes the
adsorption site for methanol molecules rather than adsorbing other acetone molecules. Fig. 7c
displays the RDFs of O_meth - H_meth and O_ace - H_meth distances for acetone-methanol
mixture in 0.8-3.0 nm split pore model at 20 kPa. The RDFs for O_meth - H_meth and O_ace
- H_meth pairs both exhibit a first peak at an intermolecular distance of 0.18 nm, confirming
the hydrogen-bonding interactions between O_ace and H_meth atoms. Moreover, the Eads of
adsorbate-adsorbate interaction systems are established by DFT-D3. The Eads for the adsorbed
acetone molecule on the pore walls to catch another acetone and methanol molecule are -18.4
kJ/mol and -36.2 kJ/mol, respectively (see Fig. 7d), which also suggests the C-OH hydrogen
bond interaction between acetone and methanol molecules is stronger than the vdW interaction
between acetone and acetone molecules.

Overall, the above results together reveal the separation mechanism of acetone-methanol
at high pressures. The entire molecular packing process in the co-adsorption of acetone-
methanol mixture can be separated into three stages. At the beginning of acetone-methanol co-
adsorption, acetone molecules will preferentially adsorb on the pore wall. Then once the first
acetone monolayer is formed on the pore walls, this will serve as the polar adsorption sites for
methanol molecules because of the stronger intermolecular interaction between acetone and
methanol (C-OH hydrogen bond interaction). The competitive adsorption of acetone and
methanol on the acetone monolayer is the main feature of this stage. Finally, methanol
molecules, with a higher packing efficiency, cluster on the acetone monolayer and gradually
fill the unoccupied pore sites by the O-HO hydrogen bonding with the increasing pressure.
Compared with the 0.8-1.6 nm micropores, the 2.0-3.0 nm mesopores have more unfilled pore
sites after adsorbing a layer of acetone molecules. This explains why mesoporous carbon has
a better methanol selectivity than microporous one.
17
5. Conclusion

In this study, the adsorption and separation performance of microporous and mesoporous
activated carbons (MIC and MEC) for acetone-methanol mixture had been investigated. The
optimal separation pore size for the acetone and methanol mixture at 293 K were first estimated
by the GCMC simulations. The 3.0 nm mesopore exhibited the largest methanol uptake
amount (19.3 mmol/cm3) and selectivity (~22.1) at 20 kPa. Then the amounts of methanol
(acetone) captured by MEC and MIC from the 50/50 mixture at high-concentration conditions
(total concentration at 202120 ppm) during the breakthrough process were 10.66 (1.60) and
6.01 (5.15) mmol/g, respectively. MEC exhibited a greater methanol/acetone selectivity (6.66)
than MIC (1.17). The experimental data corresponding to the theoretical results proved that
mesoporous carbons (with the pore size distribution range from 2.0-3.0 nm) have better
methanol selectivity at high pressure, and was more suitable for the separation of high
concentration acetone-methanol mixture gas. Finally, GCMC calculations and weak
interaction analysis suggested that different structure between acetone and methanol is the
main factor of their different interacting forms (gas-framework and gas-gas interactions),
which ultimately determined their separation process in ACs materials. Moreover, three-steps
molecular packing mode of the acetone-methanol mixture in mesopores at high pressure were
established by the RDFs and the visualized results. The 2.0-3.0 nm mesopore had more unfilled
pores space after adsorbing a layer of acetone molecules, and thus had a significant selectivity
towards methanol.

Acknowledgement

This work was supported by the National Natural Science Foundation of China (No.
21878338), the National Key Research and Development Program of China (Nos.
2019YFC0214302, 2019YFC0214303), Provincial Natural Science Foundation of Hunan
(2022JJ30015), Provincial Natural Science Foundation of Hunan (2022JJ30015) and
Environmental Protection Scientific Research Project of Hunan Province (HBKT-202102).

18
References

[1] A. Ghalandari, Z. Saadati, B.M. Goodajdar, A. Farajtabar, Solvent Effect and Preferential
Solvation Analysis of Isophthalic Acid Solubility in Acetone (1) + Water (2) and Acetic Acid
(1) + Water (2) Mixtures, J. Solution Chem. 51(9) (2022) 1148-1161.

[2] Y. Cong, C.B. Du, K. Xing, Y.C. Bian, X.X. Li, M.L. Wang, Investigation on co-solvency,
solvent effect, Hansen solubility parameter and preferential solvation of fenbufen dissolution
and models correlation, J. Mol. Liq. 348 (2022).

[3] P. Rajabzadeh, P. Ghanbarpour, E. Rahimpour, W.E. Acree, A. Jouyban, N. Aliasgharlou,


A.F. Azarbayjani, Non-electrostatic energies as a metric for prediction of deferasirox solubility
in binary solvent mixtures: Polarized Continuum Model tactic, J. Mol. Liq. 339 (2021).

[4] C.C. Chen, Y.H. Huang, J.Y. Fang, Hydrophobic deep eutectic solvents as green
absorbents for hydrophilic VOC elimination, J. Hazard. Mater. 424 (2022).

[5] C.K. Zhou, K. Zhou, H. Li, X. Xu, B.G. Liu, H.L. Li, Z. Zeng, W.W. Ma, L.Q. Li, Pressure
swing adsorption properties of activated carbon for methanol, acetone and toluene, Chem. Eng.
J. 413 (2021).

[6] L. Lyu, Q. Xie, Y.Y. Yang, R.R. Wang, W.F. Cen, S.Y. Luo, W.S. Yang, Y. Gao, Q.Q.
Xiao, P. Zou, Y. Yang, A novel CeO2 Hollow-Shell sensor constructed for high sensitivity of
acetone gas detection, Appl. Surf. Sci. 571 (2022).

[7] Y.Z. Zheng, Y.M. Lu, X.Y. Tian, H.Y. He, Y.C. Zhang, Combination of FTIR and DFT to
study the regulation law of [EMIM][OAc] on the microstructure of the acetone-methanol
azeotrope system, J. Mol. Liq. 362 (2022).

[8] Z.J. Li, S. Srebnik, Expanding carbon capture capacity: uncovering additional CO2
adsorption sites in imine-linked porous organic cages, Phys. Chem. Chem. Phys. 23 (2021)
10311-10320.

[9] A. Yz, A. Ht, B. Dw, C. Mhs, D. Cwc, C. Ysy, Predicting adsorption of micropollutants
on non-functionalized and functionalized multi-walled carbon nanotubes: Experimental study
and LFER modelling, J. Hazard. Mater. 411 (2021) 125124.

[10] M.A.H. Badsha, M. Khan, B. Wu, A. Kumar, I.M.C. Lo, Role of surface functional groups
of hydrogels in metal adsorption: From performance to mechanism, J. Hazard. Mater. 408
19
(2021) 124463.

[11] S. Salehi, M. Hosseinifard, Evaluation of CO2 and CH4 adsorption using a novel amine
modified MIL-101-derived nanoporous carbon/polysaccharides nanocomposites: Isotherms
and thermodynamics, Chem. Eng. J. 410 (1-4) (2020) 128315.

[12] B. Nfaa, A. Mj, A. Jp, A. Hj, C. Kl, D. Hoa, Flexible nanoporous activated carbon cloth
for achieving high H2, CH4, and CO2 storage capacities and selective CO2/CH4 separation,
Chem. Eng. J. 379 (2020) 122367-.

[13] X. Ma, B. Liu, Q. Wu, D. Li, R. Chen, Z. Zeng, L. Li, Specific Li+ sites in a nanoporous
carbon for enhanced light hydrocarbons storage and separation: GCMC and DFT simulations,
FUEL 288 (2021) 119647.

[14] 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.

[15] Y. Guo, Y. Guo, Z. Zeng, H.L. Li, Competitive adsorption of methanol-acetone on surface
functionalization (−COOH, −OH, −NH2 , and −SO3H): Grand canonical monte carlo and
density functional theory simulations, ACS Appl. Mater. Inter. 11(37) (2019) 34241-34250.

[16] B. Liu, L. Yu, H. Wang, X. Ma, Z. Zeng, L. Li, Experimental and molecular perspective
on VOCs adsorption and separation: Study of the surface heterogeneity and oxygen
functionalizing, Chem. Eng. J. 435 (2022) 135069.

[17] L.G. Yu, Y. Guo, H.Y. Chen, B.G. Liu, X. Xu, P. Sheng, Z. Zeng, L.Q. Li, Influence of
functional groups and pore sizes in porous carbon for methanol acetone adsorptive separation
based on molecular simulation, J Mater. Sci. 56(33) (2021) 18550-18565.

[18] Juan, José, Gutiérrez-Sevillano, Sofia, Calero, Rajamani, Krishna, Separation of benzene
from mixtures with water, methanol, ethanol, and acetone: highlighting hydrogen bonding and
molecular clustering influences in CuBTC, Phys. Chem. Chem. Phys. 17 (2015) 20114-20124.

[19] X. Ma, M. Fang, B. Liu, R. Chen, R. Shi, Q. Wu, Z. Zeng, L. Li, Urea-assisted synthesis
of biomass-based hierarchical porous carbons for the light hydrocarbons adsorption and
separation, Chem. Eng. J. 428 (2022) 130985.

[20] A. Torres-Knoop, J. Heinen, R. Krishna, D. Dubbeldam, Entropic separation of


20
styrene/ethylbenzene mixtures by exploitation of subtle differences in molecular
configurations in ordered crystalline nanoporous adsorbents, Langmuir 31(12) (2015) 3771-
3778.

[21] R. Kishan, Motkuri, K. Praveen, Thallapally, V.R. Harsha, Annapureddy, X. Liem, Dang,
Rajamani, Krishna, Separation of polar compounds using a flexible metal-organic framework,
Chem. Commun. 51 (2015) 8421-8424.

[22] P. Zhao, G. Zhang, L. Hao, A novel blended amine functionalized porous silica adsorbent
for carbon dioxide capture, Adsorption 26(5) (2020) 749-764.

[23] C.M. Simon, B. Smit, M. Haranczyk, pyIAST: Ideal adsorbed solution theory (IAST)
Python package, Comput. Phys. Commun. 200 (2016) 364-380.

[24] D. Dubbeldam, S. Calero, D.E. Ellis, R.Q. Snurr, RASPA: molecular simulation software
for adsorption and diffusion in flexible nanoporous materials, Mol. Simulat. 42(2) (2016) 81-
101.

[25] D. Dubbeldam, A. Torres-Knoop, K.S. Walton, On the inner workings of Monte Carlo
codes, Mol. Simulat. 39(14-15) (2013) 1253-1292.

[26] M.G. Martin, J.I. Siepmann, Novel configurational-bias monte carlo method for branched
molecules. Transferable potentials for phase equilibria. 2. United-atom description of
branched alkanes, J. Phys. Chem. B 103(21) (1999) 4508-4517.

[27] S.L. Mayo, B.D. Olafson, W.A. Goddard, DREIDING: A generic force field for
molecular simulations, J. Phys. Chem. 94(26) (1990) 8897-8909.

[28] K.A. Maerzke, N.E. Schultz, R.B. Ross, S. J Ilja, TraPPE-UA force field for acrylates and
Monte Carlo simulations for their mixtures with alkanes and alcohols, J. Phys. Chem. B 113(18)
(2009) 6415-25.

[29] B. Chen, J.J. Potoff, J.I. Siepmann, Monte Carlo Calculations for Alcohols and Their
Mixtures with Alkanes. Transferable Potentials for Phase Equilibria. 5. United-Atom
Description of Primary, Secondary, and Tertiary Alcohols, J.phys.chem.b 105(15) (2015)
3093-3104.

[30] A.V. Shevade, S. Jiang, K.E. Gubbins, Molecular Simulation Study of Water-Methanol
Mixtures in Activated Carbon Pores, J. Chem. Phys. 113(16) (2000) 6933-6942.
21
[31] Y. Wu, H. Chen, J. Xiao, D. Liu, Z. Liu, Y. Qian, H. Xi, Adsorptive Separation of
Methanol-Acetone on Isostructural Series of Metal-Organic Frameworks M-BTC (M = Ti, Fe,
Cu, Co, Ru, Mo): A Computational Study of Adsorption Mechanisms and Metal-Substitution
Impacts, ACS Appl. Mater. Interfaces 7(48) (2015) 26930-40.

[32] C. Xuan, B. Hou, W. Xia, Z. Peng, T. Shen, H.L. Xin, G. Zhang, D. Wang, From a ZIF-
8 polyhedron to three-dimensional nitrogen doped hierarchical porous carbon: an efficient
electrocatalyst for the oxygen reduction reaction, J Mater. Chem. A 6 (2018) 10731-10739.

[33] L. Qie, W.-M. Chen, Z.-H. Wang, Q.-G. Shao, X. Li, L.-X. Yuan, X.-L. Hu, W.-X. Zhang,
Y.-H. Huang, Nitrogen-Doped Porous Carbon Nanofiber Webs as Anodes for Lithium Ion
Batteries with a Superhigh Capacity and Rate Capability, Adv. Mater. 24(15) (2012) 2047-
2050.

[34] C. Zhou, K. Zhou, H. Li, X. Xu, B. Liu, H. Li, Z. Zeng, W. Ma, L. Li, Pressure swing
adsorption properties of activated carbon for methanol, acetone and toluene, Chem. Eng. J.
413 (2021) 127384.

[35] S. Liu, Y. Zhao, B. Zhang, H. Xia, J. Zhou, W. Xie, H. Li, Nano-micro carbon spheres
anchored on porous carbon derived from dual-biomass as high rate performance
supercapacitor electrodes, J. Power Sources 381 (2018) 116-126.

[36] L. Wang, Z. Hao, G. Cao, W. Zhang, Y. Yang, Effect of activated carbon surface
functional groups on nano-lead electrodeposition and hydrogen evolution and its applications
in lead-carbon batteries, Electrochim. Acta 186(5) (2015) 654-663.

[37] X.X. Liu, S.L. Zuo, N.N. Cui, S.S. Wang, Investigation of ammonia/steam activation for
the scalable production of high-surface area nitrogen-containing activated carbons, CARBON
191 (2022) 581-592.

[38] L. Tian, F. Chen, Multiwfn: A multifunctional wavefunction analyzer, J. Comput. Chem.


33(5) (2012) 580-592.

[39] T. Lu, Q. Chen, Van der Waals potential: An important complement to molecular
electrostatic potential in studying intermolecular interactions, ChemRxiv (2020).

[40] CP2K: An electronic structure and molecular dynamics software package -- Quickstep:
efficient and accurate electronic structure calculations, J. Chem. Phys. 152(19) (2020) 194103.

22
[41] Y. Guo, C. Su, H. Chen, L. Yu, K. Zhou, X. Xu, Z. Zeng, L. Li, Heteroatom-doped
nanoporous carbons for ethanol/benzene separation: Emphasizing the role of molecular
clustering effects, J Phys. Chem. C 125 (42) (2021) 23463-23473.

[42] Krishna, Rajamani, Diffusion in porous crystalline materials, Chem. Soc. Rev. 41(8)
(2012) 3099-3118.

Highlights

 The molecular packing steps play a decisive role in the separation processes.

 Molecular structure dictates the separation preferences of acetone and methanol.

 The 2.0-3.0 nm mesopores showed higher selectivity for methanol/acetone than the micropores.

 The self-created slit model demonstrates the three-step molecular filling process.

Graphical Abstract

23
*Credit Author Statement

Yang Guo, Liqing Li, and Zheng Zeng designed the study. Yang Guo and Changqing Su,
Hongyu Chen, Lingyun Yu, Jinxian Wang carried out experiments, simulation calculations,
and data analysis. Yang Guo drafted the manuscript. Baogen Li provided suggestions for this
study. Yang Guo, Changqing Su, Liqing Li, and Zheng Zeng revised the manuscript.

All authors read and approved the final manuscript

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:

24

You might also like