ArticlePDF Available

Large-Scale Screening and Machine Learning to Predict the Computation-Ready, Experimental Metal-Organic Frameworks for CO2 Capture from Air

Authors:

Abstract

The rising level of CO2 in the atmosphere has attracted attention in recent years. The technique of capturing CO2 from higher CO2 concentrations, such as power plants, has been widely studied, but capturing lower concentrations of CO2 directly from the air remains a challenge. This study uses high-throughput computer (Monte Carlo and molecular dynamics simulation) and machine learning (ML) to study 6013 computation-ready, experimental metal-organic frameworks (CoRE-MOFs) for CO2 adsorption and diffusion properties in the air with very low concentrations of CO2. First, the law influencing CO2 adsorption and diffusion in air is obtained as a structure-performance relationship, and then the law influencing the performance of CO2 adsorption and diffusion in air is further explored by four ML algorithms. Random forest (RF) was considered the optimal algorithm for prediction of CO2 selectivity, with an R value of 0.981, and this algorithm was further applied to analyze the relative importance of each metal-organic framework (MOF) descriptor quantitatively. Finally, 14 MOFs with the best properties were successfully screened out, and it was found that a key to capturing a low concentration CO2 from the air was the diffusion performance of CO2 in MOFs. When the pore-limiting diameter (PLD) of a MOF was closer to the CO2 dynamic diameter, this MOF could possess higher CO2 diffusion separation selectivity. This study could provide valuable guidance for the synthesis of new MOFs in experiments that capture directly low concentration CO2 from the air.
applied
sciences
Article
Large-Scale Screening and Machine Learning to
Predict the Computation-Ready, Experimental
Metal-Organic Frameworks for CO2Capture from Air
Xiaomei Deng, Wenyuan Yang, Shuhua Li, Hong Liang *, Zenan Shi * and Zhiwei Qiao *
Guangzhou Key Laboratory for New Energy and Green Catalysis, School of Chemistry and Chemical
Engineering, Guangzhou University, Guangzhou 510006, China; 1705200045@e.gzhu.edu.cn (X.D.);
2111705055@e.gzhu.edu.cn (W.Y.); lish@gzhu.edu.cn (S.L.)
*Correspondence: lhong@gzhu.edu.cn (H.L.); zenanshi@126.com (Z.S.); zqiao@gzhu.edu.cn (Z.Q.)
Received: 19 December 2019; Accepted: 10 January 2020; Published: 13 January 2020


Abstract:
The rising level of CO
2
in the atmosphere has attracted attention in recent years. The
technique of capturing CO
2
from higher CO
2
concentrations, such as power plants, has been
widely studied, but capturing lower concentrations of CO
2
directly from the air remains a
challenge. This study uses high-throughput computer (Monte Carlo and molecular dynamics
simulation) and machine learning (ML) to study 6013 computation-ready, experimental metal-organic
frameworks (CoRE-MOFs) for CO
2
adsorption and diusion properties in the air with very low
concentrations of CO
2
. First, the law influencing CO
2
adsorption and diusion in air is obtained as a
structure-performance relationship, and then the law influencing the performance of CO
2
adsorption
and diusion in air is further explored by four ML algorithms. Random forest (RF) was considered the
optimal algorithm for prediction of CO
2
selectivity, with an Rvalue of 0.981, and this algorithm was
further applied to analyze the relative importance of each metal-organic framework (MOF) descriptor
quantitatively. Finally, 14 MOFs with the best properties were successfully screened out, and it was
found that a key to capturing a low concentration CO
2
from the air was the diusion performance
of CO
2
in MOFs. When the pore-limiting diameter (PLD) of a MOF was closer to the CO
2
dynamic
diameter, this MOF could possess higher CO
2
diusion separation selectivity. This study could
provide valuable guidance for the synthesis of new MOFs in experiments that capture directly low
concentration CO2from the air.
Keywords:
CO
2
capture; Monte Carlo; machine learning; metal–organic framework; adsorption;
diusion
1. Introduction
It is well known that the amount of CO
2
discharged into the atmosphere increases with the
rapid development of industry and population growth. In addition, deforestation, the large amount
of CO
2
and other gases generated by the burning of fossil fuels such as coal, oil, and natural gas
directly discharged into the atmosphere, and the emission of limestone roasting to produce cement
have resulted in global carbon dioxide emissions increasing by 3.8% [
1
]. All of the above factors have
aggravated carbon dioxide emissions, thereby increasing the urgency of counteracting the greenhouse
eect and its associated global warming. The Kyoto Protocol and the Paris Agreement aim to control
greenhouse gas emissions under the United Nations Framework Convention on Climate Change
(UNFCCC), in which CO
2
is listed as a major greenhouse gas that needs to be mitigated or recycled [
2
].
The greenhouse gases include more than CO
2
, however; in fact, the global warming potentials of CH
4
and N
2
O are 25 times and 298 times that of CO
2
, respectively. Nevertheless, due to its relatively large
emission levels, CO
2
accounts for approximately 55% of the total greenhouse gas contribution [
3
,
4
].
Appl. Sci. 2020,10, 569; doi:10.3390/app10020569 www.mdpi.com/journal/applsci
Appl. Sci. 2020,10, 569 2 of 13
Thus, it is obvious that the adsorption and separation of carbon dioxide from the air is particularly
important. In addition, the successful capture of CO
2
could have multifaceted practical values: first,
oil recovery could be improved through appropriate reservoir engineering; second, the captured CO
2
could be used to produce industrial chemicals, including concrete, paint, and fertilizer; third, the CO
2
in the atmosphere could be captured and combined with hydrogen for direct synthesis into liquid
hydrocarbons, which could then be utilized in fuel synthesis and supply, including gasoline and diesel.
The use of raw materials can reduce the proportion of fossil energy to further control CO
2
emissions,
ultimately achieving carbon neutrality or even net negative carbon emissions [5].
Recently, carbon engineering has developed a series of capture technologies that remove carbon
dioxide directly from the air. Carbon dioxide can be removed from the atmosphere using biological,
chemical, or physical processes [
6
]. These methods have certain limitations, however. For example,
biological processes are very economical, but they are usually very slow and ineective. As for
chemical processes, the waste of carbon resources and volatilization of organic solvents during these
actions lead to further environmental pollution, equipment corrosion, and complex post-treatment
issues. The traditional technique for separating carbon dioxide is solvent washing, such as the use of
an alcohol amine solution [
7
10
]. Although this conventional method can reduce the concentration of
carbon dioxide in the air, it is extremely expensive, the solvent is dicult to regenerate, the operation
is complicated, and it consumes a great deal of energy [
11
]. In fact, the energy consumption of solvent
washing is 3 to 4 times that of CO
2
captured from exhaust gas [
12
]. Given these drawbacks, there
is an urgent need to find a more ecient, convenient, and energy-saving technique to replace the
traditional carbon dioxide capture method. Adsorption separation is a potential technique. It is
not only inexpensive, but also simple in terms of operation and equipment, and relatively low in
energy consumption when the adsorbent is regenerated (the regeneration process of adsorbents is
to desorb the adsorbed substances). Conventional adsorbents, however, including activated carbon,
zeolite, silica gel, and metal oxides have poor scavenging eects on carbon dioxide in the air due to
inferior separation selectivity and regeneration diculty. For example, silica gel, which has amorphous
properties, does not have a continuous uniform porous structure and exhibits unfavorable diusion
properties [
11
]. Therefore, the development of a new type of adsorbent is imperative. In recent years,
studies have shown that the use of metal-organic frameworks (MOFs) to adsorb and separate carbon
dioxide can not only make up for the shortcomings of the above adsorbents, but also feature the
advantages of high selectivity and being non-polluting. The MOF is an organic–inorganic hybrid
material with intramolecular pores formed by the self-assembly of organic ligands and inorganic metal
ions or clusters by coordination bonds [
13
]. Compared with common adsorbents, MOFs exhibit many
advantages such as various structures and properties, large specific surface area, high porosity, and
structural control. Therefore, they are widely used in gas adsorption [
11
] and separation [
14
19
],
as well as general materials in processes including storage [
20
], optics [
21
], catalysis [
22
25
], and
drug delivery [
26
,
27
]. To date, thousands of MOFs have been synthesized, some of which have
been utilized in the attempt to capture CO
2
from the air. Peng et al. [
28
] designed and synthesized
2 incorporated MOFs to study their stability and ability to capture CO
2
from the air. Liu et al. [
11
]
used an amine-functionalized MOF and an ultra-microporous MOF to capture CO
2
directly from
the air, and further investigated the performance of CO
2
capture and the reproducibility of MOFs
under humid conditions. Osama et al. [
29
] synthesized an isomorphic MOF SIFSIX-3-Cu with uniform
adsorption sites for capturing CO
2
from the air. Since CO
2
capture from the air has a very high
selectivity of MOF, when the traditional approach is used to screen MOFs for the best-performing
candidates, it not only consumes a great deal of manpower and material resources, but also has an
extended study period and causes pollution to a certain extent. With the continuous advancement and
development of computers, molecular simulation is playing an increasingly important role in the field
of materials science [
30
]. Some studies have used high-throughput molecular simulation calculation
methods to screen large numbers of MOFs in a database, thereby successfully screening MOFs with
high selectivity and high working capacity based on dierent target performance requirements. For
Appl. Sci. 2020,10, 569 3 of 13
example, Wilmer et al. adsorbed pure carbon dioxide, nitrogen, and methane using more than 130,000
hypothetical MOFs, and proposed a relationship between structural properties (pore size, volume, and
surface area) and chemical functions, as well as evaluation criteria for the separation of carbon dioxide
from adsorbents [
31
]. In the presence of nickel dilution, Watanabe et al. combined pore size analysis
with classical simulation to screen 1163 MOFs as membrane materials for CO
2
/N
2
separation [
32
].
Lin et al. screened hundreds of thousands of theoretically predicted zeolites and zeolite MOFs and
identified a number of potential materials for capturing carbon dioxide [
33
]. Based on 105 MOFs, Wu
et al. proposed the relationship of CO
2
/N
2
adsorption selectivity with porosity and the isosteric heat of
adsorption [
34
]. Fernandez et al. [
35
] used advanced machine-learning (ML) algorithms to quickly
identify 292,050 hypothetical high-performance MOFs for pure CO
2
adsorption (0.15 bar and 1 bar).
These screening studies, however, were aimed at capturing high concentrations of CO
2
. Given that the
concentration of CO
2
in the atmosphere is comparatively low relative to the concentrations of natural
gas and other components, it is undoubtedly a challenge to discover ecient MOF materials that can
directly capture CO2from the air.
To date, given that there have been 6013 MOFs reported, finding the appropriate MOFs for a
specific system in such a large database is undoubtedly a daunting task. This study focused on the
aforementioned MOF simulation of the adsorption and diusion performances of CO
2
, N
2
, and O
2
in infinite dilutions in order to identify materials with excellent performance in terms of both static
adsorption and kinetic adsorption. The influencing factors aecting the adsorption and diusion
of CO
2
were obtained by univariate analysis. Next, multivariate analyses, namely 4 ML algorithms
(back propagation neural network (BPNN), decision tree (DT), random forest (RF), and support
vector machine (SVM)), were explored in depth. Finally, we adopted the optimal algorithm model.
The parameters aecting CO
2
selectivity were predicted, and 14 types of MOFs with the same diusion
selectivity and adsorption selectivity were selected.
2. Materials and Methods
2.1. Molecular Model
In this work, we used molecular simulation to screen the capability of 6013 computation-ready,
experimental MOFs (CORE-MOFs version 2) [
36
] to capture CO
2
from the air. Their crystal structures
were derived from the Cambridge Crystallographic Data Centre (CCDC), and their parameters were
compiled and verified by Chung et al. [
37
]. We removed all solvent and ligand molecules prior to
running the simulation. Each MOF used 5 structural parameters, namely, volumetric surface area
(VSA), largest cavity diameter (LCD), pore-limiting diameter (PLD), porosity
φ
, density
ρ
, and an
energy parameter: heat of adsorption. Both LCD and PLD were calculated using the Zeo++ software
package [
38
]. The VSA and
φ
were calculated using the N
2
of 3.64 Å and He of 2.58 Å as probes in
the RASPA software package [
39
]. If VSA is close to or equal to 0, this indicates that the MOF cannot
accommodate N
2
molecules [
40
]. We used NVT-Monte Carlo (NVT-MC) simulation, where N is the
number of particles, V is the volume of the system, and T is the temperature of the system. The Q
st
of
each gas was calculated in an infinite dilution state.
The force field parameters for the 3 gas components CO
2
/N
2
/O
2
were from the transferable
potentials for phase equilibria (TraPPE) force field [
41
] and are listed in Table S2 The CO
2
molecule has
a C-O bond length of 1.16 Å and a bond angle
OCO of 180
. N
2
is considered as a 3-point model,
and the bond length of N-N is 1.10 Å. Oxygen is also a 3-point atom, and the O-O bond length is 1.21 Å.
The models of 3 gases are shown in Figure S1, The atomic charge of MOF was estimated using the MOF
electrostatic-potential-optimized charge scheme (MEPO-Qeq) method [
42
], which accurately evaluated
electrostatic interactions. Due to the advantages of the MEPO-Qeq method with fast and accurate,
it is widely used in various systems of adsorption-MOF [
43
45
]. The Lennard–Jones (LJ) electrostatic
parameters were obtained from the universal force field (UFF) [
46
] and are listed in Table S1 Data from
previous studies had shown that the UFF–TraPPE force field combination could accurately predict
Appl. Sci. 2020,10, 569 4 of 13
the adsorption and diusion behaviors of these 3 gases in MOFs [
40
,
47
,
48
]. The Lorentz–Berthelot
combination rule was used to calculate the cross-LJ parameters.
2.2. Screening Methods
In MOFs, the values of Henry’s constant Kand the diusion coecient Dof CO
2
, N
2
, and O
2
were estimated using Monte Carlo (MC) and molecular dynamics (MD) simulations with the same
set, respectively. In principle, a single gas molecule should be added to an MOF to simulate infinite
dilution, while in reality, we added 30 gas molecular models to each MOF, ignoring the force between
the gas models, thus being equivalent to the independent simulation of each gas molecule. Ultimately,
the simulation results of the 30 independent molecules were statistically averaged. Throughout the
simulation, the MOF frame was assumed to be rigid and the simulation elements were extended
to at least 24 Å along the three-dimensional periodic boundary conditions. A 12 Å spherical cuto
with long-range correction was used to calculate the LJ interaction, while the Ewald sum was used to
calculate the electrostatic interaction. In each MOF, the MC simulation ran 100,000 cycles, with the
first 50,000 used for balancing and the last 50,000 used for overall averaging. Each cycle consisted of n
trials (n: number of adsorbed molecules), including translation, rotation, regeneration, and exchange
(exchange movement, including insertion and deletion). In the MD simulation, the 30 gas molecules
had an MD duration of 10 ns at each MOF, and 5 ns was ultimately selected for statistical averaging.
After the sampling analysis of dozens of MOFs, it was found that further increases of cycle time and
MD duration had little eect on the simulation results. All MCs and MDs were simulated using the
RASPA software package [39].
3. Results and Discussion
3.1. Univariate Analysis
In order to investigate the relationship of CO
2
adsorption and diusion properties in N
2
+O
2
with
the MOF structure during static adsorption and kinetic adsorption, we first analyzed the relationship
among adsorption selectivity S
ads
, diusion selective S
di
, and the LCD of CO
2
/N
2
+O
2
, as shown
in Figure 1. Obviously, most MOFs with large adsorption selectivity and diusion selectivity have
relatively small LCDs. Figure 1a indicates that when the LCD is 2.8–6.5 Å, the adsorption selectivity
of CO
2
/N
2
+O
2
decreases, and when the LCD is >15 Å, the adsorption selectivity gradually becomes
stable, tending to 5, as depicted by the red line in Figure 1This is because CO
2
has a strong quadrupole
moment, and even in infinitely large pores it is preferentially adsorbed compared to N
2
and O
2
.
The trend of gas separation is consistent with the trends of previous reports [
49
,
50
]. Figure 1b presents
the relationship between S
di
and LCD. Similar to S
ads
, the larger diusion selectivity (S
di
>1) only
occurs in the LCDs ranging from 2.4–5 Å, since the kinetic diameter of CO
2
is less than the kinetic
diameters of O
2
and N
2
(the kinetic diameters of CO
2
, O
2
, and N
2
are 3.3, 3.46, and 3.64 Å, respectively).
When the LCD of an MOF is small, the CO
2
molecules with smaller diameters diuse faster, so the
diusion selectivity S
di(CO2/N2+O2)
is larger. As the LCD increases, the diusion selectivity gradually
decreases. When the LCD is >15 Å, the diusion selectivity tends to be stable and fluctuates at around
0.36. Comparison of Figure 1a,b reveals that the adsorption selectivity is generally >1, while it is rare
for the diusion selectivity to be >1. Because CO
2
has a strong quadrupole distance, it has a strong
interaction force with MOF molecules, thus hindering the diusion of CO
2
and resulting in a slower
diusion rate, which may be even smaller than the diusion rates of N2and O2.
Figure 1c,d show the relationships of S
ads
and S
di
to the PLD, respectively. Comparing the panels
in Figure 1reveals that the PLD and LCD display the same trend in their relationships to the S
ads
and
S
di
of CO
2
/N
2
+O
2
. Larger S
ads
and S
di
values appear when the LCD and PLD are small, and S
ads
and S
di
both decrease with increasing PLD or LCD, eventually tending toward stability. Therefore,
there is a greater possibility of finding MOFs with simultaneously high S
ads
and S
di
among MOFs
with small PLDs and LCDs.
Appl. Sci. 2020,10, 569 5 of 13
Appl. Sci. 2020, 10, x FOR PEER REVIEW 5 of 13
5 1015202530
10-3
100
103
106
109
S
ads(CO2/(N2+O2))
LCD (Å)
(a)
5 1015202530
10-3
10-2
10-1
100
101
102
S
diff(CO2/(N2+O2))
LCD (Å)
(b)
51015
10-3
100
103
106
109
1012
S
ads(CO2/(N2+O2))
PLD ( )
(c)
51015
10-2
10-1
100
101
102
S
diff(CO2/(N2+O2))
PLD
(
Å
(d)
Figure 1. The relationships of (a) Sads (CO2/N2+O2) and (b) Sdiff (CO2/N2+O2) with (c) largest cavity diameter
(LCD) and (d) pore-limiting diameter (PLD).
Figure 2a shows that Sads (CO2/N2+O2) increases monotonically with increasing Qst, indicating that
Qst may be the main parameter during the adsorption process. Since the concentration of CO2 in the
atmosphere is low, it is close to the infinite dilution state. Hence, the selectivity is strongly
dependent on the isosteric heat of adsorption of CO2 in the infinite dilution state. The larger Sdiff
(CO2/N2+O2) in Figure 2b occurs when the VSA is close to zero. As the VSA continues to increase, Sdiff
(CO2/N2+O2) gradually decreases, and eventually stabilizes. This is because when the VSA is close to
zero, the MOF molecule either cannot pass any or only passes a small amount of CO2 molecules.
When the VSA is large, all the gas molecules can pass through the MOF molecule. Therefore, the
separation of CO2 cannot be achieved, i.e., the diffusion selectivity is substantially unchanged.
Figure S11b,c indicate the relationship of adsorption selectivity with porosity and VSA, respectively.
It can be observed that both of these parameters exert weak influences on adsorption selectivity.
0 20 40 60 80 100
10-1
102
105
108
1011
1014
S
ads(CO2/(N2+O2))
Q
st (kJ/mol)
(a)
0 1000 2000 3000
10-3
10-2
10-1
100
101
102
S
diff(CO2/(N2+O2))
VSA
(
m2/cm3
)
(b)
Figure 2. The relationships between (a) Sads (CO2/N2+O2) and Qst, (b) Sdiff (CO2/N2+O2) and volumetric surface
area (VSA).
Figure 1.
The relationships of (
a
) S
ads (CO2/N2+O2)
and (
b
) S
di(CO2/N2+O2)
with (
c
) largest cavity diameter
(LCD) and (d) pore-limiting diameter (PLD).
Figure 2a shows that S
ads (CO2/N2+O2)
increases monotonically with increasing Q
st
, indicating that
Q
st
may be the main parameter during the adsorption process. Since the concentration of CO
2
in the
atmosphere is low, it is close to the infinite dilution state. Hence, the selectivity is strongly dependent
on the isosteric heat of adsorption of CO
2
in the infinite dilution state. The larger S
di(CO2/N2+O2)
in
Figure 2b occurs when the VSA is close to zero. As the VSA continues to increase, S
di(CO2/N2+O2)
gradually decreases, and eventually stabilizes. This is because when the VSA is close to zero, the MOF
molecule either cannot pass any or only passes a small amount of CO
2
molecules. When the VSA is
large, all the gas molecules can pass through the MOF molecule. Therefore, the separation of CO
2
cannot be achieved, i.e., the diusion selectivity is substantially unchanged. Figure S11b,c indicate the
relationship of adsorption selectivity with porosity and VSA, respectively. It can be observed that both
of these parameters exert weak influences on adsorption selectivity.
Appl. Sci. 2020, 10, x FOR PEER REVIEW 5 of 13
5 1015202530
10-3
100
103
106
109
S
ads(CO2/(N2+O2))
LCD (Å)
(a)
5 1015202530
10-3
10-2
10-1
100
101
102
S
diff(CO2/(N2+O2))
LCD (Å)
(b)
51015
10-3
100
103
106
109
1012
S
ads(CO2/(N2+O2))
PLD ( )
(c)
51015
10-2
10-1
100
101
102
S
diff(CO2/(N2+O2))
PLD
(
Å
(d)
Figure 1. The relationships of (a) Sads (CO2/N2+O2) and (b) Sdiff (CO2/N2+O2) with (c) largest cavity diameter
(LCD) and (d) pore-limiting diameter (PLD).
Figure 2a shows that Sads (CO2/N2+O2) increases monotonically with increasing Qst, indicating that
Qst may be the main parameter during the adsorption process. Since the concentration of CO2 in the
atmosphere is low, it is close to the infinite dilution state. Hence, the selectivity is strongly
dependent on the isosteric heat of adsorption of CO2 in the infinite dilution state. The larger Sdiff
(CO2/N2+O2) in Figure 2b occurs when the VSA is close to zero. As the VSA continues to increase, Sdiff
(CO2/N2+O2) gradually decreases, and eventually stabilizes. This is because when the VSA is close to
zero, the MOF molecule either cannot pass any or only passes a small amount of CO2 molecules.
When the VSA is large, all the gas molecules can pass through the MOF molecule. Therefore, the
separation of CO2 cannot be achieved, i.e., the diffusion selectivity is substantially unchanged.
Figure S11b,c indicate the relationship of adsorption selectivity with porosity and VSA, respectively.
It can be observed that both of these parameters exert weak influences on adsorption selectivity.
0 20 40 60 80 100
10-1
102
105
108
1011
1014
S
ads(CO2/(N2+O2))
Q
st (kJ/mol)
(a)
0 1000 2000 3000
10-3
10-2
10-1
100
101
102
S
diff(CO2/(N2+O2))
VSA
(
m2/cm3
)
(b)
Figure 2. The relationships between (a) Sads (CO2/N2+O2) and Qst, (b) Sdiff (CO2/N2+O2) and volumetric surface
area (VSA).
Figure 2.
The relationships between (
a
) S
ads (CO2/N2+O2)
and Q
st
, (
b
) S
di(CO2/N2+O2)
and volumetric
surface area (VSA).
Appl. Sci. 2020,10, 569 6 of 13
In addition to adsorption and diusion selectivity, the Henry coecient of CO
2
reflects the
adsorption performance of CO
2
in the infinite dilution state, helping to explain the capture performance
of MOFs for air with very low CO
2
concentration. Figure 3a clearly shows the tendency of K
N2
to
change with enthalpy. When the porosity
φ
is small, the MOF has no space due to the limited pore
volume, and only a small amount of N
2
can be adsorbed; therefore, K
N2
is small. When
φ
is in the
range of 0–0.29, K
N2
increases significantly with increasing
φ
. When
φ
>0.29, K
N2
slows down and
gradually stabilizes with increasing
φ
. Figure 3b compares the Henry coecients of the 3 gases. It can
be seen that the trends of the Henry coecients of N
2
and O
2
are almost the same; however, CO
2
is dierent. First, in most MOFs, the Henry coecient values of CO
2
are basically larger than the
Henry coecient values of N
2
and O
2
. Second, when LCD >20 Å, the K
CO2
value tends to be level,
and eventually stabilizes. The Henry coecient of CO
2
is still higher than the coecients of N
2
and
O
2
, which also leads to MOF selectivity >1 when the LCD is infinite, as seen in Figure 1a Finally, it can
be observed that only a few MOFs can be identified for which the CO
2
Henry coecient can be >10
1
mmol/g/Pa. Observing these MOF structures reveals that most have smaller or open metal sites. The
above univariate analysis can only determine the relationship between individual parameters and
performance. Q
st
, PLD and LCD are considered to have dramatic impacts on adsorption selectivity and
diusion selectivity, but their variable influences cannot be analyzed quantitatively. We will further
utilize 4 types of ML algorithms to obtain additional information about structure-performance.
Appl. Sci. 2020, 10, x FOR PEER REVIEW 6 of 13
In addition to adsorption and diffusion selectivity, the Henry coefficient of CO2 reflects the
adsorption performance of CO2 in the infinite dilution state, helping to explain the capture
performance of MOFs for air with very low CO2 concentration. Figure 3a clearly shows the tendency
of KN2 to change with enthalpy. When the porosity ϕ is small, the MOF has no space due to the
limited pore volume, and only a small amount of N2 can be adsorbed; therefore, KN2 is small. When ϕ
is in the range of 0–0.29, KN2 increases significantly with increasing ϕ. When ϕ > 0.29, KN2 slows down
and gradually stabilizes with increasing ϕ. Figure 3b compares the Henry coefficients of the 3 gases.
It can be seen that the trends of the Henry coefficients of N2 and O2 are almost the same; however,
CO2 is different. First, in most MOFs, the Henry coefficient values of CO2 are basically larger than the
Henry coefficient values of N2 and O2. Second, when LCD >20 Å, the KCO2 value tends to be level, and
eventually stabilizes. The Henry coefficient of CO2 is still higher than the coefficients of N2 and O2,
which also leads to MOF selectivity >1 when the LCD is infinite, as seen in Figure 1a Finally, it can be
observed that only a few MOFs can be identified for which the CO2 Henry coefficient can be >101
mmol/g/Pa. Observing these MOF structures reveals that most have smaller or open metal sites. The
above univariate analysis can only determine the relationship between individual parameters and
performance. Qst, PLD and LCD are considered to have dramatic impacts on adsorption selectivity
and diffusion selectivity, but their variable influences cannot be analyzed quantitatively. We will
further utilize 4 types of ML algorithms to obtain additional information about
structure-performance.
0.0 0.2 0.4 0.6 0.8 1.0
10-17
10-14
10-11
10-8
10-5
10-2
K
N2
(mmol/g/Pa)
φ
(a)
5 1015202530
10-13
10-10
10-7
10-4
10-1
102
105 K CO2
K N2
K O2
K
LCD (Å)
(b)
Figure 3. Henry coefficient KN2 versus (a) ϕ and (b) LCD.
3.2. Machine Learning
Currently, machine learning has been used to predict the performance of materials and to filter
high-performance materials from large databases [51]. Aiming to discover a better machine
prediction method suitable for this system, we individually compared the simulations of the 4 ML
algorithms commonly used in big data analysis, i.e., the BPNN, DT, RF, and SVM. Among them,
BPNN is a kind of forward signal propagation with error back propagation in which the gradient
descent algorithm continuously adjusts the weight and threshold until the error is less than a set
threshold. Some parameters of BPNN were set: the training function is Levenberg–Marquardt, the
transfer function is a hyperbolic tangent sigmoid transfer function, and the performance evaluation
function is the mean square error (MSE). The number of hidden layer neurons was 18, the maximum
number of training was 1000, the training required an accuracy of 0.001, and the learning rate was
0.01. DT is a traditional method for data classification and screening. Under the condition that the
probability of occurrence takes place in various situations, probability analysis is employed to
analyze data with the dendritic model to obtain the expected values. The random forest algorithm is
composed of multiple decision trees. The setting parameters of DT were: standard CART
(classification and regression tree) used to select the best split predictor at each node. The criteria of
splitting and pruning are the MSE function. After optimizing and pruning, the minimum number of
branch node observations was 10, the minimum number of leaf node observations was 4, and the
maximal number of decision splits was 1. RF uses the method of randomly selecting split attribute
Figure 3. Henry coecient KN2 versus (a)φand (b) LCD.
3.2. Machine Learning
Currently, machine learning has been used to predict the performance of materials and to filter
high-performance materials from large databases [
51
]. Aiming to discover a better machine prediction
method suitable for this system, we individually compared the simulations of the 4 ML algorithms
commonly used in big data analysis, i.e., the BPNN, DT, RF, and SVM. Among them, BPNN is a kind
of forward signal propagation with error back propagation in which the gradient descent algorithm
continuously adjusts the weight and threshold until the error is less than a set threshold. Some
parameters of BPNN were set: the training function is Levenberg–Marquardt, the transfer function
is a hyperbolic tangent sigmoid transfer function, and the performance evaluation function is the
mean square error (MSE). The number of hidden layer neurons was 18, the maximum number of
training was 1000, the training required an accuracy of 0.001, and the learning rate was 0.01. DT is a
traditional method for data classification and screening. Under the condition that the probability of
occurrence takes place in various situations, probability analysis is employed to analyze data with the
dendritic model to obtain the expected values. The random forest algorithm is composed of multiple
decision trees. The setting parameters of DT were: standard CART (classification and regression
tree) used to select the best split predictor at each node. The criteria of splitting and pruning are the
MSE function. After optimizing and pruning, the minimum number of branch node observations
was 10, the minimum number of leaf node observations was 4, and the maximal number of decision
splits was 1. RF uses the method of randomly selecting split attribute sets to construct a decision tree.
The parameters for RF were set as: number of trees 200, minimum leaf size 10. The number of variables
Appl. Sci. 2020,10, 569 7 of 13
randomly selected in the variable subset of the node split in each tree was 2. SVM is an algorithm for
binary classification of data through supervised learning, and employs mathematical transformation
methods to divide data with a certain centralized structure into rules. We chose the support vector
machine regression model of Statistics and Machine Learning Toolbox in MATLAB 2016b to predict,
where the kernel function is radial basis function (Gaussian), the kernel scale parameter is set as “auto”,
and the loss function is epsilon-insensitive. The box constraint (also called the penalty coecient, C)
was 0.0567, and the half width of epsilon-insensitive band (
ε
) was set as 0.0057. In the radial basis
kernel function, Gamma =1/(2
σ2
), where
σ
is the parameter of the kernel function, which can aect the
complexity of the SVM regression algorithm. In our study, the value of gamma was 7.125. The solver of
convex quadratic programming is sequential minimal optimization (SMO). Before training and testing,
we first processed the data set out-of-order, and then randomly divide it into training and testing
sets based on a ratio of 7:3. More detailed descriptions of ML algorithms are listed in the supporting
information, and the corresponding diagrams of each algorithm are shown in Figures S2–S5.
We used BPNN, RF, DT, and SVM to predict the adsorption selectivity, and took the logarithm of
the adsorption selectivity in order to reduce the dierences associated with the varying magnitudes
of the data. The 4 ML for predicting the correlation coecient Rvalue of the adsorption selectivity
are listed in Table 1The results of the testing and training are shown in Figure 4and Figure S12.
The distribution trends of the points in Figure 4a–d are all straight lines inclined upward. The dierent
colors from top to bottom in the figure represent an increase in the number of points, and most of
the points are concentrated on the diagonal, indicating that the prediction results are good. Figure 4
reveals that RF has the highest correlation coecient value (0.982), while the support vector machine
algorithm has the lowest (0.886). Thus, the prediction accuracy obtained by the RF algorithm is
the highest. Therefore, among the 4 ML algorithms, the structure-performance relationship of RF
on adsorption selectivity obtains more information, and the prediction results are the best. RF has
good generalization ability and strong model learning ability, and this type of ML is suitable for the
system. To verify the accuracy of four ML algorithms, we performed 5 repeated predictions, listed in
Table S3. The repeated prediction results do not vary significantly, confirming that the RF algorithm is
a suitable model. Because RF introduces two kinds of randomness (sampling randomness and feature
randomness), it has strong generalization ability. In previous studies, random forest algorithms also
exhibited the best prediction results [
52
,
53
]. Whether the overfitting of the model is an important
issue. We used combinations of dierent descriptors and 5 times 5-fold cross-validation to verify
the RF model. The results showed that the selected model was not overfitting. During the material
screening process, the relative importance of parameters may aect the ultimate screening results.
We selected the best RF algorithm to predict the relative importance of each descriptor. The relative
importance percentages are shown in Figure 5and Table S7. We used mean squared error (MSE) to
evaluate the relative importance of the 6 descriptors; the greater the percentage of relative importance
of the resulting descriptors, the higher the relative contribution of the specific descriptor. According to
the results presented in Figure 5, the percentage of Q
st
is the largest, thus indicating that Q
st
exerts the
greatest impact on the adsorption selectivity. The relative importance of the MOF descriptors to the
adsorption selectivity is Q
st
>
ρ
>LCD >VSA
φ
>PLD. The importances of VSA and
φ
are very
close, indicating that the eects of two descriptors on adsorption selectivity are roughly equal. From a
material science point of view, the larger the
φ
, the larger the VSA. This may be the reason why the
eects of these parameters are essentially the same.
Table 1.
The 4 ML algorithms for predicting the correlation coefficient Rvalue of the adsorption selectivity.
ML Algorithms R Value
Train Test
BPNN 0.982 0.979
RF 0.994 0.981
DT 0.985 0.969
SVM 0.915 0.886
Appl. Sci. 2020,10, 569 8 of 13
Appl. Sci. 2020, 10, x FOR PEER REVIEW 8 of 13
Figure 4. The test results of four machine-learning (ML) algorithms predicted versus simulated
Sads(CO2/N2+O2) using ρ, ϕ, VSA, LCD, PLD and heat of adsorption. (a) back propagation neural network
(BPNN), (b) random forest (RF), (c) decision tree (DT), (d) support vector machine (SVM). The color
of point represents the number of metal-organic frameworks (MOFs), and the unit of number is a
base-10 logarithm of MOF numbers.
Table 1. The 4 ML algorithms for predicting the correlation coefficient R value of the adsorption
selectivity.
ML Algorithms R Value
Train Test
BPNN 0.982 0.979
RF 0.994 0.981
DT 0.985 0.969
SVM 0.915 0.886
Figure 4.
The test results of four machine-learning (ML) algorithms predicted versus simulated
S
ads(CO2/N2+O2)
using
ρ
,
φ
, VSA, LCD, PLD and heat of adsorption. (
a
) back propagation neural
network (BPNN), (
b
) random forest (RF), (
c
) decision tree (DT), (
d
) support vector machine (SVM).
The color of point represents the number of metal-organic frameworks (MOFs), and the unit of number
is a base-10 logarithm of MOF numbers.
Appl. Sci. 2020, 10, x FOR PEER REVIEW 9 of 13
LCD VSA PLD
0
10
20
30
40
50
60
φ
Relative importance(%)
MOF Descriptors
ρ
Q
st
Figure 5. Predicted by the Random Forest, the relative importance of the six descriptors for
adsorption selectivity.
4. Best Metal-Organic Frameworks (MOFs)
We selected 5 limiting conditions for Sads (CO2/N2+O2) and Sdiff (CO2/N2+O2), and chose 14 optimal MOFs
from the 6013 MOFs, as listed in Table 2 of the 14 materials, HIQPEE exhibited the largest Sdiff, which
was as much as 62.27 Å. NORGOS displayed the largest Sads, which also corresponded to its
maximum heat of adsorption. In comparison, the optimal MOF selected by this study is also more
selective at higher Qst (4712.33) under the same conditions than that predicted by Wu et al. (433) [30]
at Qst = 47.8 kJ/mol and Ravichandar Babarao et al. (500) [54]. It was discovered that diffusion
selectivity is generally lower than adsorption selectivity. The diffusion of CO2 is the key property in
determining the performance of MOFs for low concentrations of CO2 during the kinetic adsorption
process. For these 14 MOFs, the LCD, ϕ, and PLD ranges of the six descriptors also corresponded to
those in the previous univariate analysis. Especially for PLDs, the optimal range of 2.66–3.64 Å only
spans 1 Å, which is also very close to the kinetic diameter of the 3 gases. In such strictly restricted
channels, only CO2 molecules can enter and be adsorbed, greatly increasing the probability of CO2
being captured at low concentrations. Therefore, the analysis of the optimal MOF revealed that a
PLD with a kinetic diameter close to CO2 is a key condition for good CO2 diffusion performance,
further resulting in the excellent performance of the MOF in capturing CO2 from the air, and thus
providing effective guidance for the design and synthesis of new MOFs.
Table 2. Best computation-ready, experimental metal-organic frameworks (CoRE-MOFs).
No CSD Code
a LCD b ϕ VSA c
(m2/cm3) PLD d (Å) Ρ (kg/m3) Qst_CO2
(kJ/mol) SdiffCO2/(N2+O2) SadsCO2/(N2+O2)
1 REYCEF 3.75 0.08 0 2.83 1646.40 26.02 8.21 6.25
2 JAHNEM 3.50 0.05 0 2.66 1714.62 27.01 6.18 6.77
3 HOJLEY 3.63 0.14 0 2.83 1410.99 32.65 5.00 7.45
4 KASPOL 4.04 0.19 16.26 2.93 1665.95 34.05 36.55 7.68
5 XUNJOG 3.25 0.07 0 2.67 1737.75 26.41 5.31 11.12
6 HIQPII 3.87 0.15 9.46 3.12 1472.40 35.20 55.67 11.24
7 YUBFUX 4.58 0.16 98.56 3.64 1786.84 30.52 5.79 11.51
8 HIQPEE 3.84 0.15 7.68 3.12 1440.14 35.31 62.27 12.39
9 FEJKEM 3.46 0.09 0.33 3.09 2132.02 30.00 15.01 15.72
10 FALQIU 5.08 0.12 82.58 3.14 1977.64 34.40 5.38 22.97
11 FALQOA 6.06 0.12 83.34 3.07 2004.41 35.03 6.76 24.10
12 TOXNAX 3.80 0.28 4.42 2.83 1346.52 42.65 5.02 25.15
13 OFIWIK 4.20 0.05 5.99 3.14 1866.86 39.64 27.42 25.66
14 NORGOS 4.95 0.13 45.40 3.43 1728.90 51.67 12.92 4712.33
a CSD Code is the code of MOFs in the Cambridge Structure Database; b LCD: largest cavity diameter; c VSA: volumetric
surface area; d PLD: pore-limiting diameter.
Figure 5.
Predicted by the Random Forest, the relative importance of the six descriptors for
adsorption selectivity.
Appl. Sci. 2020,10, 569 9 of 13
4. Best Metal-Organic Frameworks (MOFs)
We selected 5 limiting conditions for S
ads (CO2/N2+O2)
and S
di(CO2/N2+O2),
and chose 14 optimal
MOFs from the 6013 MOFs, as listed in Table 2of the 14 materials, HIQPEE exhibited the largest
S
di
, which was as much as 62.27 Å. NORGOS displayed the largest S
ads
, which also corresponded
to its maximum heat of adsorption. In comparison, the optimal MOF selected by this study is also
more selective at higher Q
st
(4712.33) under the same conditions than that predicted by Wu et al.
(433) [
30
] at Q
st
=47.8 kJ/mol and Ravichandar Babarao et al. (500) [
54
]. It was discovered that diusion
selectivity is generally lower than adsorption selectivity. The diusion of CO
2
is the key property in
determining the performance of MOFs for low concentrations of CO
2
during the kinetic adsorption
process. For these 14 MOFs, the LCD,
φ
, and PLD ranges of the six descriptors also corresponded to
those in the previous univariate analysis. Especially for PLDs, the optimal range of 2.66–3.64 Å only
spans 1 Å, which is also very close to the kinetic diameter of the 3 gases. In such strictly restricted
channels, only CO
2
molecules can enter and be adsorbed, greatly increasing the probability of CO
2
being captured at low concentrations. Therefore, the analysis of the optimal MOF revealed that a PLD
with a kinetic diameter close to CO
2
is a key condition for good CO
2
diusion performance, further
resulting in the excellent performance of the MOF in capturing CO
2
from the air, and thus providing
eective guidance for the design and synthesis of new MOFs.
Table 2. Best computation-ready, experimental metal-organic frameworks (CoRE-MOFs).
No CSD
Code aLCD bφVSA c
(m2/cm3)PLD d
(Å)
P
(kg/m3)
Qst_CO2
(kJ/mol) SdiCO2/(N2+O2) SadsCO2/(N2+O2)
1 REYCEF 3.75 0.08 0 2.83 1646.40 26.02 8.21 6.25
2 JAHNEM 3.50 0.05 0 2.66 1714.62 27.01 6.18 6.77
3 HOJLEY 3.63 0.14 0 2.83 1410.99 32.65 5.00 7.45
4 KASPOL 4.04 0.19 16.26 2.93 1665.95 34.05 36.55 7.68
5 XUNJOG 3.25 0.07 0 2.67 1737.75 26.41 5.31 11.12
6 HIQPII 3.87 0.15 9.46 3.12 1472.40 35.20 55.67 11.24
7 YUBFUX 4.58 0.16 98.56 3.64 1786.84 30.52 5.79 11.51
8 HIQPEE 3.84 0.15 7.68 3.12 1440.14 35.31 62.27 12.39
9 FEJKEM 3.46 0.09 0.33 3.09 2132.02 30.00 15.01 15.72
10 FALQIU 5.08 0.12 82.58 3.14 1977.64 34.40 5.38 22.97
11 FALQOA 6.06 0.12 83.34 3.07 2004.41 35.03 6.76 24.10
12 TOXNAX 3.80 0.28 4.42 2.83 1346.52 42.65 5.02 25.15
13 OFIWIK 4.20 0.05 5.99 3.14 1866.86 39.64 27.42 25.66
14 NORGOS 4.95 0.13 45.40 3.43 1728.90 51.67 12.92 4712.33
a
CSD Code is the code of MOFs in the Cambridge Structure Database;
b
LCD: largest cavity diameter;
c
VSA:
volumetric surface area; dPLD: pore-limiting diameter.
5. Conclusions
Firstly, we simulated the adsorption and diusion properties of CO
2
, N
2
, and O
2
in 6013
CoRE-MOFs using high-throughput MC +MD. Then, we investigated the correlation among adsorption
selectivity and diusion selectivity for CO
2
and MOF descriptors by the univariate analysis. The Q
st
and PLD were considered to be the most important for S
ads (CO2/N2+O2)
and S
di(CO2/N2+O2)
, respectively.
In conjunction with multivariate analysis, a comparison of 4 ML algorithms revealed that the RF had
the best prediction results for adsorption selectivity, with an Rvalue of 0.982. This indicated that
the RF method was the most suitable for the predictions of the capture of low CO
2
concentrations in
MOF. The relative importance analysis of the RF algorithm quantitatively indicated that the relative
importance of the MOF descriptors on adsorption selectivity is Q
st
>
ρ
>LCD >VSA
φ
>PLD.
It was also confirmed that Q
st
is the most important parameter, while the VSA and
φ
are relatively
less important. Through this high-throughput screening, 14 types of MOFs with optimal adsorption
selectivity and diusion selectivity were obtained. After comparison, it was found that their adsorption
selectivity was generally higher than their diusion selectivity. The diusion separation performance of
CO
2
is the key property in determining the performance of MOFs on low concentrations of CO
2
during
Appl. Sci. 2020,10, 569 10 of 13
the kinetic adsorption process. This study provides experimental guidance for the determination
of MOFs that eectively capture CO
2
from the air, and indicates that advanced ML algorithms can
accelerate the research and development of new materials.
Supplementary Materials:
The following are available online at http://www.mdpi.com/2076-3417/10/2/569/s1:
Figure S1: Models of N
2
, O
2
and CO
2
, Figure S2: Back propagation neural network, Figure S3: Decision tree,
Figure S4: Random forest, Figure S5: Support vector machine, Figure S6: Diusion coecient D
i
versus (a) LCD,
(b–d) Q
st
(D
i
: the diusion coecient of dierent gases, (a,b) CO
2
, (c) N
2
, (d) O
2
), Figure S7: Henry coecient
K
CO2
versus (a)
φ
, (b) PLD, (c) Q
st
and (d) VSA, Figure S8: Henry coecient K
N2
versus (a)
φ
, (b) PLD, (c) Q
st
and (d)
VSA, Figure S9: Henry coecient K
O2
versus (a)
φ
, (b) PLD, (c) Q
st
and (d) VSA, Figure S10: Diusion selectivity
S
di(CO2/(N2+O2))
versus (a) LCD, (b)
φ
, (c) PLD and (d) VSA, Figure S11: Adsorption selectivity S
ads (CO2/N2+O2)
versus (a) LCD, (b)
φ
, (c) PLD and (d) VSA, Figure S11: The training results of four machine learning predicted
versus MC simulated log
10
(S
ads (CO2/N2+O2)
) using density, void fraction, volumetric surface area, density, LCD,
PLD and heat of adsorption. (a) BPNN, (b) RF, (c) DT and (d) SVM. The color of point represents the number of
MOF, and the unit of mumber is a base-10 logarithm of MOF numbers, Table S1: Lennard-Jones parameters of
MOFs, Table S2: Lennard-Jones parameters and charges of adsorbates, Table S3: The training and testing Rvalues
of adsorption selectivity using repeat 5 time-four ML. Table S4 Prediction using RF models with dierent descriptor
combinations. Table S5 Prediction using repeat 5 times-RF models with dierent descriptor combinations. Table
S6 The results of predicted RF with ktimes k-fold cross validation. Table S7: Predicted by the RF the relative
importance of the six descriptors for adsorption selectivity. Formula S1: S
ads (CO2/N2+O2)
indicates the adsorption
selectivity of CO
2
/N
2
+O
2
;K
i
represents the Henry coecient of component i(CO
2
, N
2
and O
2
), Formula S2:
S
di(CO2/(N2+O2))
represents the diusion selectivity of CO
2
/N
2
+O
2
;D
i
represents the diusion coecient of
component i.
Author Contributions:
X.D., Z.S., H.L. and Z.Q. conceived the idea. Z.Q. calculated all the materials’ structural
parameters and obtained valid data about the structure descriptors and performance. X.D., W.Y. and Z.S. analyzed
the relationship between structure descriptors and performance. X.D. and S.L. used univariate analysis to obtain
the influence law of aecting CO
2
adsorption and diusion in air and Z.S. used ML algorithms to predict the MOF
performance. X.D. and Z.S. wrote the original draft. H.L. and Z.Q. wrote the manuscript with contributions from
all authors. All authors have read and agreed to the published version of the manuscript.
Funding:
This research was funded by the National Natural Science Foundation of China (Nos. 21978058,
21676094, and 21576058).
Acknowledgments:
We gratefully thank the National Natural Science Foundation of China (Nos. 21978058,
21676094, and 21576058) for financial support.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Canadell, J.G.; Le Quere, C.; Raupach, M.R.; Field, C.B.; Buitenhuis, E.T.; Ciais, P.; Conway, T.J.; Gillett, N.P.;
Houghton, R.A.; Marland, G. Contributions to accelerating atmospheric CO
2
growth from economic activity,
carbon intensity, and eciency of natural sinks. Proc. Natl. Acad. Sci. USA
2007
,104, 18866–18870. [CrossRef]
[PubMed]
2.
Zhang, C.; Zeng, G.; Huang, D.; Lai, C.; Chen, M.; Cheng, M.; Tang, W.; Tang, L.; Dong, H.; Huang, B.; et al.
Biochar for environmental management: Mitigating greenhouse gas emissions, contaminant treatment, and
potential negative impacts. Chem. Eng. J. 2019,373, 902–922. [CrossRef]
3.
Fan, R.; Chen, C.-L.; Lin, J.-Y.; Tzeng, J.-H.; Huang, C.-P.; Dong, C.; Huang, C.P. Adsorption characteristics of
ammonium ion onto hydrous biochars in dilute aqueous solutions. Bioresour. Technol.
2019
,272, 465–472.
[CrossRef] [PubMed]
4.
Fang, G.; Liu, C.; Wang, Y.; Dionysiou, D.D.; Zhou, D. Photogeneration of reactive oxygen species from
biochar suspension for diethyl phthalate degradation. Appl. Catal. B Environ. 2017,214, 34–45. [CrossRef]
5. Available online: https://carbonengineering.com/(accessed on 10 July 2019).
6. Nibleus, K.; Lundin, R. Climate Change and Mitigation. Ambio 2010,39, 11–17. [CrossRef]
7.
Boyd, P.G.; Chidambaram, A.; Garc
í
a-D
í
ez, E.; Ireland, C.P.; Da, T.D.; Bounds, R.; Gładysiak, A.;
Schouwink, P.; Moosavi, S.M.; Maroto-Valer, M.M.; et al. Data-driven design of metal–organic frameworks
for wet flue gas CO2capture. Nature 2019,576, 253–256. [CrossRef]
8.
Faig, R.W.; Popp, T.M.O.; Fracaroli, A.M.; Kapustin, E.A.; Kalmutzki, M.J.; Altamimi, R.M.; Fathieh, F.;
Reimer, J.A.; Yaghi, O.M. The Chemistry of CO
2
Capture in an Amine-Functionalized Metal-Organic
Framework under Dry and Humid Conditions. J. Am. Chem. Soc. 2017,139, 12125–12128. [CrossRef]
Appl. Sci. 2020,10, 569 11 of 13
9.
Haszeldine, R.S. Carbon Capture and Storage: How Green Can Black Be? Science
2009
,325, 1647–1652.
[CrossRef]
10.
McDonald, T.M.; Mason, J.A.; Kong, X.; Bloch, E.D.; Gygi, D.; Dani,A.; Crocell
à
, V.; Giordanino, F.; Odoh, S.O.;
Drisdell, W.S.; et al. Cooperative insertion of CO
2
in diamine-appended metal-organic frameworks. Nature
2015,519, 303–308. [CrossRef]
11.
Liu, J.; Wei, Y.; Zhao, Y. Trace Carbon Dioxide Capture by Metal-Organic Frameworks. ACS Sustain.
Chem. Eng. 2019,7, 82–93. [CrossRef]
12.
Zhao, R.; Liu, L.; Zhao, L.; Deng, S.; Li, S.; Zhang, Y.; Li, H. Thermodynamic exploration of temperature
vacuum swing adsorption for direct air capture of carbon dioxide in buildings. Energy Convers. Manag.
2019
,
183, 418–426. [CrossRef]
13.
Batten, S.R.; Champness, N.R.; Chen, X.-M.; Garcia-Martinez, J.; Kitagawa, S.; Ohrstrom, L.; O’Keee, M.;
Suh, M.P.; Reedijk, J. Terminology of metal-organic frameworks and coordination polymers (IUPAC
Recommendations 2013). Pure Appl. Chem. 2013,85, 1715–1724. [CrossRef]
14.
Murray, L.J.; Dinca, M.; Long, J.R. Hydrogen storage in metal-organic frameworks. Chem. Soc. Rev.
2009
,38,
1294–1314. [CrossRef] [PubMed]
15.
Sculley, J.; Yuan, D.; Zhou, H.-C. The current status of hydrogen storage in metal-organic frameworks-updated.
Energy Environ. Sci. 2011,4, 2721–2735. [CrossRef]
16.
Li, J.-R.; Kuppler, R.J.; Zhou, H.-C. Selective gas adsorption and separation in metal-organic frameworks.
Chem. Soc. Rev. 2009,38, 1477–1504. [CrossRef] [PubMed]
17.
Verma, S.; Mishra, A.K.; Kumar, J. The Many Facets of Adenine: Coordination, Crystal Patterns, and Catalysis.
Acc. Chem. Res. 2010,43, 79–91. [CrossRef]
18.
Li, J.-R.; Sculley, J.; Zhou, H.-C. Metal-Organic Frameworks for Separations. Chem. Rev.
2012
,112, 869–932.
[CrossRef]
19.
Bae, Y.-S.; Snurr, R.Q. Development and Evaluation of Porous Materials for Carbon Dioxide Separation and
Capture. Angew. Chem. Int. Ed. 2011,50, 11586–11596. [CrossRef]
20.
Wu, X.-J.; Zhao, P.; Fang, J.-M.; Wang, J.; Liu, B.-S.; Cai, W.-Q. Simulation on the Hydrogen Storage Properties
of New Doping Porous Aromatic Frameworksl. Acta Phys. Chim. Sin. 2014,30, 2043–2054. [CrossRef]
21.
Wu, P.; He, C.; Wang, J.; Peng, X.; Li, X.; An, Y.; Duan, C. Photoactive Chiral Metal-Organic Frameworks
for Light-Driven Asymmetric alpha-Alkylation of Aldehydes. J. Am. Chem. Soc.
2012
,134, 14991–14999.
[CrossRef]
22.
Farrusseng, D.; Aguado, S.; Pinel, C. Metal-Organic Frameworks: Opportunities for Catalysis. Angew. Chem.
Int. Ed. 2009,48, 7502–7513. [CrossRef] [PubMed]
23.
Ma, L.; Abney, C.; Lin, W. Enantioselective catalysis with homochiral metal-organic frameworks.
Chem. Soc. Rev. 2009,38, 1248–1256. [CrossRef] [PubMed]
24.
Lee, J.; Farha, O.K.; Roberts, J.; Scheidt, K.A.; Nguyen, S.T.; Hupp, J.T. Metal-organic framework materials as
catalysts. Chem. Soc. Rev. 2009,38, 1450–1459. [CrossRef] [PubMed]
25.
Farha, O.K.; Shultz, A.M.; Sarjeant, A.A.; Nguyen, S.T.; Hupp, J.T. Active-Site-Accessible, Porphyrinic
Metal-Organic Framework Materials. J. Am. Chem. Soc. 2011,133, 5652–5655. [CrossRef] [PubMed]
26.
Della Rocca, J.; Liu, D.; Lin, W. Nanoscale Metal-Organic Frameworks for Biomedical Imaging and Drug
Delivery. Acc. Chem. Res. 2011,44, 957–968. [CrossRef]
27.
Bernini, M.C.; Fairen-Jimenez, D.; Pasinetti, M.; Ramirez-Pastor, A.J.; Snurr, R.Q. Screening of bio-compatible
metal-organic frameworks as potential drug carriers using Monte Carlo simulations. J. Mater. Chem. B
2014
,
2, 766–774. [CrossRef]
28.
Peng, Y.-W.; Wu, R.-J.; Liu, M.; Yao, S.; Geng, A.-F.; Zhang, Z.-M. Nitrogen Coordination to Dramatically
Enhance the Stability of In-MOF for Selectively Capturing CO2from a CO2/N2Mixture. Cryst. Growth Des.
2019,19, 1322–1328. [CrossRef]
29.
Shekhah, O.; Belmabkhout, Y.; Chen, Z.; Guillerm, V.; Cairns, A.; Adil, K.; Eddaoudi, M. Made-to-order
metal-organic frameworks for trace carbon dioxide removal and air capture. Nat. Commun.
2014
,5.
[CrossRef]
30.
Jain, A.; Shyue Ping, O.; Hautier, G.; Chen, W.; Richards, W.D.; Dacek, S.; Cholia, S.; Gunter, D.; Skinner, D.;
Ceder, G.; et al. Commentary: The Materials Project: A materials genome approach to accelerating materials
innovation. APL Mater. 2013,1. [CrossRef]
Appl. Sci. 2020,10, 569 12 of 13
31.
Furukawa, H.; Cordova, K.E.; O’Keee, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic
Frameworks. Science 2013,341, 1230444. [CrossRef]
32.
Watanabe, T.; Sholl, D.S. Accelerating Applications of Metal-Organic Frameworks for Gas Adsorption and
Separation by Computational Screening of Materials. Langmuir
2012
,28, 14114–14128. [CrossRef] [PubMed]
33.
Lin, L.-C.; Berger, A.H.; Martin, R.L.; Kim, J.; Swisher, J.A.; Jariwala, K.; Rycroft, C.H.; Bhown, A.S.;
Deem, M.W.; Haranczyk, M.; et al. In silico screening of carbon-capture materials. Nat. Mater.
2012
,11,
633–641. [CrossRef] [PubMed]
34.
Wu, D.; Yang, Q.; Zhong, C.; Liu, D.; Huang, H.; Zhang, W.; Maurin, G. Revealing the Structure-Property
Relationships of Metal-Organic Frameworks for CO
2
Capture from Flue Gas. Langmuir
2012
,28, 12094–12099.
[CrossRef] [PubMed]
35.
Fernandez, M.; Boyd, P.G.; Da, T.D.; Aghaji, M.Z.; Woo, T.K. Rapid and Accurate Machine Learning
Recognition of High Performing Metal Organic Frameworks for CO
2
Capture. J. Phys. Chem. Lett.
2014
,5,
3056–3060. [CrossRef] [PubMed]
36.
Available online: https://github.com/gregchung/gregchung.github.io/blob/master/CoRE-MOFs/structure-
doi-CoRE-MOFsV2.0.csv (accessed on 15 January 2019).
37.
Chung, Y.G.; Camp, J.; Haranczyk, M.; Sikora, B.J.; Bury, W.; Krungleviciute, V.; Yildirim, T.; Farha, O.K.;
Sholl, D.S.; Snurr, R.Q. Computation-Ready, Experimental Metal-Organic Frameworks: A Tool to Enable
High-Throughput Screening of Nanoporous Crystals. Chem. Mater. 2014,26, 6185–6192. [CrossRef]
38.
Willems, T.F.; Rycroft, C.; Kazi, M.; Meza, J.C.; Haranczyk, M. Algorithms and tools for high-throughput
geometry-based analysis of crystalline porous materials. Microporous Mesoporous Mater.
2012
,149, 134–141.
[CrossRef]
39.
Dubbeldam, D.; Calero, S.; Ellis, D.E.; Snurr, R.Q. RASPA: Molecular simulation software for adsorption and
diusion in flexible nanoporous materials. Mol. Simul. 2016,42, 81–101. [CrossRef]
40.
Yang, W.; Liang, H.; Peng, F.; Liu, Z.; Liu, J.; Qiao, Z. Computational Screening of Metal-Organic Framework
Membranes for the Separation of 15 Gas Mixtures. Nanomaterials 2019,9, 467. [CrossRef]
41.
Poto, J.J.; Siepmann, J.I. Vapor-liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen.
AIChE J. 2001,47, 1676–1682. [CrossRef]
42.
Kadantsev, E.S.; Boyd, P.G.; Da, T.D.; Woo, T.K. Fast and Accurate Electrostatics in Metal Organic
Frameworks with a Robust Charge Equilibration Parameterization for High-Throughput Virtual Screening
of Gas Adsorption. J. Phys. Chem. Lett. 2013,4, 3056–3061. [CrossRef]
43.
Shi, Z.; Liang, H.; Yang, W.; Liu, J.; Liu, Z.; Qiao, Z. Machine learning and in silico discovery of metal-organic
frameworks: Methanol as a working fluid in adsorption-driven heat pumps and chillers. Chem. Eng. Sci.
2020,214, 115430. [CrossRef]
44.
Qiao, Z.; Xu, Q.; Jiang, J. Computational screening of hydrophobic metal-organic frameworks for the
separation of H2S and CO2 from natural gas. J. Mater. Chem. A 2018,6, 18898–18905. [CrossRef]
45.
Bian, L.; Li, W.; Wei, Z.; Liu, X.; Li, S. Formaldehyde Adsorption Performance of Selected Metal-Organic
Frameworks from High-throughput Computational Screening. Acta Chim. Sin.
2018
,76, 303–310. [CrossRef]
46.
Rappe, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard, W.A.; Ski, W.M. UFF, a full periodic table force field
for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc.
1992
,114, 10024–10035.
[CrossRef]
47.
Qiao, Z.; Xu, Q.; Jiang, J. High-throughput computational screening of metal-organic framework membranes
for upgrading of natural gas. J. Membr. Sci. 2018,551, 47–54. [CrossRef]
48.
Babarao, R.; Jiang, J. Diusion and separation of CO
2
and CH
4
in silicalite, C
168
schwarzite, and IRMOF-1:
A comparative study from molecular dynamics simulation. Langmuir 2008,24, 5474–5484. [CrossRef]
49.
Qiao, Z.; Zhang, K.; Jiang, J. In silico screening of 4764 computation-ready, experimental metal-organic
frameworks for CO2separation. J. Mater. Chem. A 2016,4, 2105–2114. [CrossRef]
50.
Wilmer, C.E.; Farha, O.K.; Bae, Y.-S.; Hupp, J.T.; Snurr, R.Q. Structure-property relationships of porous
materials for carbon dioxide separation and capture. Energy Environ. Sci. 2012,5, 9849–9856. [CrossRef]
51.
Takahashi, K.; Tanaka, Y. Materials informatics: A journey towards material design and synthesis. Dalton Trans.
2016,45, 10497–10499. [CrossRef]
52.
Pardakhti, M.; Moharreri, E.; Wanik, D.; Suib, S.L.; Srivastava, R. Machine Learning Using Combined
Structural and Chemical Descriptors for Prediction of Methane Adsorption Performance of Metal Organic
Frameworks (MOFs). ACS Comb. Sci. 2017,19, 640–645. [CrossRef]
Appl. Sci. 2020,10, 569 13 of 13
53.
Wu, X.; Xiang, S.; Su, J.; Cai, W. Understanding Quantitative Relationship between Methane Storage
Capacities and Characteristic Properties of Metal-Organic Frameworks Based on Machine Learning. J. Phys.
Chem. C 2019,123, 8550–8559. [CrossRef]
54.
Babarao, R.; Jiang, J. Unprecedentedly High Selective Adsorption of Gas Mixtures in rho Zeolite-like
Metal-Organic Framework: A Molecular Simulation Study. J. Am. Chem. Soc.
2009
,131, 11417–11425.
[CrossRef]
©
2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
... [1,2] Metal organic frameworks (MOF) due to having properties such as high thermal stabilities, large surface areas (>8000 m2 g-1), low densities (0.2 g cm-3), and wide range of pore sizes (3-100 Å) are widely used in the absorption and separation of all types of gases. [3] Conducting experiments on all these MOFs on a laboratory scale is time-consuming, expensive and impossible. Therefore, it is essential to develop models for the prediction of CO2 storage capacity of MOFs to find the best candidates. ...
... Therefore, it is essential to develop models for the prediction of CO2 storage capacity of MOFs to find the best candidates. [3] In this work, we use ensemble methods that include multiple learning algorithms or models to produce one optimal predictive model. six machine learning-based methods are used, namely Random forest, XGBoost, LightGBM, GradientBoostingRegressor (GBDT), VotingRegressor, and BaggingRegressor. ...
Conference Paper
Metal organic frameworks (MOFs) have been extensively studied for carbon capture to combat global warming. The use of molecular simulation for adsorbent screening is computationally expensive and therefore prohibitive to materials discovery. in this work, the use of machine learning (ML) algorithms for the simulation of CO2 adsorption capacities of several tens of thousands of hypothetical MOFs are investigated at various thermodynamic conditions using ensemble methods.
Article
Environmental pollution has become a serious issue due to the rapid development of urbanization, industrialization, and vehicle traffic. Notably, fossil fuel combustion significantly contributes to rising atmospheric CO2 levels. To address this problem, various carbon capture and storage (CCS) technologies have been developed, aiming to reduce CO2 emissions and mitigate their impact on climate change. Absorption using aqueous amines has long been recognized as a method for removing diluted CO2 from gas streams, but it comes with drawbacks such as high energy intensity and corrosion issues. The use of solid adsorbents, however, is now being seriously considered as a potential alternative, offering a possibly less energy-intensive separation method. The primary focus of this study is to outline the recent development of advanced solid adsorbents, including zeolites, carbon-based materials, MOFs, COFs, boron nitride, magnetic nanoparticles, and mesoporous silica, along with their CO2 uptake behavior. In CO2 capture procedures, selecting the appropriate adsorbent is crucial, yet it's not a straightforward task to determine the most promising sorbent beforehand due to various factors affecting performance and economy. In recent years, machine learning (ML) algorithms, particularly artificial neural networks (ANN) and convolutional neural networks (CNN) have emerged as valuable tools for predicting physical properties, expediting the selection of optimal adsorbents for CO2 capture, optimizing synthesis conditions of adsorbents, and understanding advantageous variables for gas-solid interaction. The secondary objective of this review is to establish a correlation between recent advancements and potential future breakthroughs in the domain of machine learning-assisted CO2 adsorbents. In summary, this review aims to provide a comprehensive guideline for selecting tailored solid adsorbent materials according to recently reported research to achieve high-performance CO2 capture. By exploring various materials, their properties, and the mechanisms that influence their effectiveness, this review intends to facilitate informed decisions and innovative solutions for CO2 adsorbents.
Article
Direct air capture (DAC) of CO2 with porous adsorbents such as metal−organic frameworks (MOFs) has the potential to aid large-scale decarbonization. Previous screening of MOFs for DAC relied on empirical force fields and ignored adsorbed H2O and MOF deformation. We performed quantum chemistry calculations overcoming these restrictions for thousands of MOFs. The resulting data enable efficient descriptions using machine learning.
Article
The adoption of carbon capture technology in coal-fired power plants is expected to play a pivotal role in the energy transition. This study conducted consequential life cycle assessments (CLCAs) of coal-fired power generation in the United States using policy-level accounting. Monoethanolamine (MEA)-based and Mg-MOF-74-based carbon capture have been introduced, with a comparative analysis conducted on the emissions reduction potential of these two materials through their respective mechanisms of absorption and adsorption. The results indicate that carbon capture based on MEA or Mg-MOF-74 can significantly reduce emissions from coal-fired power generation, decreasing from 779.5 Mt CO2e to 50.1 Mt CO2e and 61.1 Mt CO2e in 2050, respectively. The introduction of ultra-supercritical power plants and carbon capture reduced direct emissions from 92% to 51%. MEA outperforms Mg-MOF-74 slightly, with lower emissions due to solvents and cleaning processes. Deviations in Mg-MOF-74's adsorption capacity and degradation rate could lead to 4%-6% model outcome variations. It is also concluded that the stability of MEA's marginal emissions depends on a steady expansion of existing production capacity, while the marginal emissions of Mg-MOF-74 are anticipated to remain unchanged. This study emphasizes carbon capture's potential but stresses the need for prompt implementation and comprehensive assessments before deployment decisions.
Article
Full-text available
For hydrogen-powered vehicles, the efficiency cost brought about by the current industry choices of hydrogen storage methods greatly reduces the system’s overall efficiency. The physisorption of hydrogen fuel onto metal–organic frameworks (MOFs) is a promising alternative storage method due to their large surface areas and exceptional tunability. However, the massive selection of MOFs poses a challenge for the efficient screening of top-performing MOF structures that are capable of meeting target hydrogen uptakes. This study examined the performance of 13 machine learning (ML) models in the prediction of the gravimetric and volumetric hydrogen uptakes of real MOF structures for comparison with simulated and experimental results. Among the 13 models studied, 12 models gave an R2 greater than 0.95 in the prediction of both the gravimetric and the volumetric uptakes in MOFs. In addition, this study introduces a 4-20-1 ANN model that predicts the bulk, shear, and Young’s moduli for the MOFs. The machine learning models with high R2 can be used in choosing MOFs for hydrogen storage.
Article
Full-text available
Extensive trial and error in the variable space is the main cause of low efficiency and high cost in material development. The experimental tasks can be reduced significantly in the case that the variable space is narrowed down by reliable computer simulations. Because of their numerous variables in material design, however, the variable space is still too large to be accessed thoroughly even with a computational approach. High‐throughput computations (HTC) make it possible to complete a material screening in a large space by replacing the conventionally manual and sequential operations with automatic, robust, and concurrent streamlines. The efficiency of HTC, which is one of the pillars of materials genome engineering, has been verified in many studies, but its applications are still limited by demanding computational costs. Introduction of data mining and artificial intelligence into HTC has become an effective approach to solve the problem. In the past years, many studies have focused on the development and application of HTC and data combined approaches, which is considered as a new paradigm in computational materials science. This review focuses on the main advances in the field of data‐assisted HTC for material research and development and provides our outlook on its future development.
Article
Full-text available
The question we pose in this study is to what extent the ranking of metal organic frameworks (MOFs) for adsorption-based carbon capture, and the selection of top performers identified in...
Article
Full-text available
Limiting the increase of CO2 in the atmosphere is one of the largest challenges of our generation¹. Because carbon capture and storage is one of the few viable technologies that can mitigate current CO2 emissions², much effort is focused on developing solid adsorbents that can efficiently capture CO2 from flue gases emitted from anthropogenic sources³. One class of materials that has attracted considerable interest in this context is metal–organic frameworks (MOFs), in which the careful combination of organic ligands with metal-ion nodes can, in principle, give rise to innumerable structurally and chemically distinct nanoporous MOFs. However, many MOFs that are optimized for the separation of CO2 from nitrogen4–7 do not perform well when using realistic flue gas that contains water, because water competes with CO2 for the same adsorption sites and thereby causes the materials to lose their selectivity. Although flue gases can be dried, this renders the capture process prohibitively expensive8,9. Here we show that data mining of a computational screening library of over 300,000 MOFs can identify different classes of strong CO2-binding sites—which we term ‘adsorbaphores’—that endow MOFs with CO2/N2 selectivity that persists in wet flue gases. We subsequently synthesized two water-stable MOFs containing the most hydrophobic adsorbaphore, and found that their carbon-capture performance is not affected by water and outperforms that of some commercial materials. Testing the performance of these MOFs in an industrial setting and consideration of the full capture process—including the targeted CO2 sink, such as geological storage or serving as a carbon source for the chemical industry—will be necessary to identify the optimal separation material.
Article
Full-text available
The Monte Carlo and molecular dynamics simulations are employed to screen the separation performance of 6013 computation-ready, experimental metal–organic framework membranes (CoRE-MOFMs) for 15 binary gas mixtures. After the univariate analysis, principal component analysis is used to reduce 44 performance metrics of 15 mixtures to a 10-dimension set. Then, four machine learning algorithms (decision tree, random forest, support vector machine, and back propagation neural network) are combined with k times repeated k-fold cross-validation to predict and analyze the relationships between six structural feature descriptors and 10 principal components. Based on the linear correlation value R and the root mean square error predicted by the machine learning algorithm, the random forest algorithm is the most suitable for the prediction of the separation performance of CoRE-MOFMs. One descriptor, pore limiting diameter, possesses the highest weight importance for each principal component index. Finally, the 30 best CoRE-MOFMs for each binary gas mixture are screened out. The high-throughput computational screening and the microanalysis of high-dimensional performance metrics can provide guidance for experimental research through the relationships between the multi-structure variables and multi-performance variables.
Article
With the increasingly serious energy-consumption, the adsorption-driven heat pumps/chillers were paid more attention. To select the promising adsorbents and suitable working fluids in this technology, a high-throughput computational screening of 6013 computation-ready experimental metal organic frameworks (MOFs) is employed to identity the best candidates for methanol-MOF pairs in adsorption-driven heat pumps and chillers. The highest working capacity (ΔW) and coefficient of performance (COP) under operating condition of heat pump are 512.86 mg/g and 1.83, respectively. The structure-performance relationships are derived for ΔW and COP with 6 MOF descriptors, the saturated ΔW is, for the first time, defined for the improvement of heat pump efficiency, and 10 best MOFs are identified. Then four machine learning algorithms are applied to predict the relative importance of each MOF descriptor. The microscopic insights obtained from our bottom-up approach provide the guidelines for the design of new MOFs in adsorption-driven heat pumps and chillers.
Article
With increased interest in scientific investigations and large-scale applications of lignocellulosic biomass derived biochar for environmental management, a comprehensive understanding of its function in the fragile earth environment is urgently needed. The study on the relationship between biochar properties and potential applications requires continuing research. This review provides new insights into the state-of-the-art accomplishments in the utilization of biochar in environmental management and covers three perspectives: firstly, mitigation of greenhouse gas (GHG) emissions, such as sequestration of CO 2 and CH 4 in global carbon pools and mitigation of N 2 O emissions; secondly, pollution control, including adsorptive removal and reactive removal of inorganic and organic contaminants; thirdly, potential negative aspects of biochar applications, including contaminations originated from biochar, negative alterations to soil properties and soil biota, negative impacts of biochar on GHG emissions and negative impacts of biochar migration. From a unique and comprehensive environmental perspective, this article aims to provide a critical review of updated knowledge on both positive and negative impacts of biochar for environmental management, based on an exponentially increased number of publications on the topic over the past decade.
Article
Metal organic frameworks (MOFs) are one category of the emerging porous materials, which are promising competitors applied in gas storage and separation due to their high porosity and high surface area. It is still time-consuming to search the optimal materials for methane storage from a large number of candidates by traditional methods such as molecular simulations and quantum mechanics. Recently, machine-learning (ML) algorithms were gradually used to accelerate the discovery of high-performance MOFs. In this work, Henry's coefficient besides other characteristic parameters were computed and appended into the previously reported dataset of hypothetical metal organic frameworks (hMOFs) for methane storage. The new dataset with 37 features and 130,397 samples was then randomly split into train set and test set in the ratio of 7:3, which were applied for ML training and testing with three different algorithms, including support vector machine (SVM), random forest regression (RFR) and gradient boosting regression tree (GBRT). The results indicate that GBRT model demonstrates the best generalization ability to predict non-trained dataset, while RFR model results in the best predictive power in the training set. The analysis of feature importance from machine learning algorithms confirms that high generalization ability of GBRT model is attributed to the model extracting more information from wider range of features. RFR model results in the highest prediction accuracy with Pearson correlation coefficient (r2) of 0.9984 and root mean square error (RMSE) of 3.93 in the training set of absolute gravimetric uptakes. GBRT model results in the highest prediction accuracy with r2 of 0.9908 and RMSE of 9.40 in the test set of absolute gravimetric uptakes, which is of the highest prediction accuracy among the up-to-date reports. According to volumetric capacities for methane storage, the optimal hMOFs exhibit ϕ of 0.65~0.88, LCD of ~7.5 Å and VSA of ~2250 m2 cm-3 etc.
Article
Abrupt climate change such as the loss of Arctic sea-ice area urgently needs negative emissions technologies. The potential application of direct air capture of carbon dioxide from indoor air and outdoor air in closed buildings or crowded places has been discussed in this paper. From the aspects of carbon reduction and indoor comfort, the ventilation system integrating a capture device is of great value in practical use. For ultra-dilute carbon dioxide sources, many traditional separation processes have no cost advantages, but adsorption technologies such as temperature vacuum swing adsorption is one of suitable methods. Thermodynamic exploration has been investigated regarding minimum separation work and second-law efficiency at various concentrations in the air. The influence of concentration, adsorption temperature, desorption temperature and desorption pressure on the energy efficiency has also been evaluated. Results show that the minimum separation work for the level of 400 ppm is approximately 20 kJ/mol. The optimal second-law efficiencies are 44.57%, 37.55% and 31.60%, respectively for 3000 ppm, 2000 ppm and 1000 ppm. It means that a high energy-efficiency capture device in buildings merits attention in the exploration of the possibility of approaching negative carbon buildings.
Article
Two In-based metal-organic frameworks (MOFs) are design and synthesized by reaction of 3-phenylpyridine polycarboxylic ligands and In3+ cations, (In3L14)•(CH3NH2CH3)3 (L1 = 5-(3’,5’-dicarboxylphenyl) nicotic acid, 1) and [In2L22(CH3COO)2]•[(CH3)2NH2][(CH3)3NH]•H2O (L2 = 3-(2,4-dicarboxylphenyl)-6- carboxylpyridine, 2). Single crystal X-ray diffraction analysis indicates that the In3+ centers in 1 are all coordinated by eight oxygen atoms from four carboxylates, resulting in a 3D framework with two different channels A and B with the size of 14.4 × 8.75 Å2 and 7.24 × 6.7 Å2, respectively. In 2, each In3+ center was coordinated by six carboxylate oxygen atoms and a pyridine N atom, forming a 2D layer-like structure. These 2D layers are further fused together via electrostatic interactions into a 3D supermolecular framework, showing two kinds of channels A´and B´ with the pore sizes of 5.25 × 9.69 Å2 and 4.29 × 8.66 Å2, respectively. In this work, the coordination of N to In3+ center in 2 was firstly achieved by regulating carboxyl sites in 3-phenylpyridine, thus increasing the stability of In-based MOF. Detail structural study reveals that the formation of In-N bonds promotes the conversion of 3D framework of 1 into supermolecular structure of 2. However, the supermolecular structure in 2 is more stable in air and methanol than that of the 3D framework in 1, confirming the important role of the In-N bonds in enhancing the framework stability. Thus, 2 could be used as stable adsorbent to selective capture CO2 from CO2/N2 mixture, while 1 broke down sharply in air and the commonly used exchanged solvent methanol.
Article
Climate deterioration is closely related to the CO2 concentration in atmosphere, which is considered one of the major environmental challenges we are facing today. It is urgent to take immediate actions to prevent further climate change. In comparison with post-combustion CO2 capture technologies from flue gas, trace CO2 capture directly from air is still a challenge but very important for both CO2 control in atmosphere and air condition control in confined space. This review highlights recent research advances in the use of metal-organic frameworks (MOFs) for trace CO2 capture, with the emphasis on employing amine functionalized MOFs and ultra-microporous MOFs. Herein, the reported strategies to enhance CO2 adsorption capacity and selectivity by MOFs are categorized into three main directions, including the developments of MOFs with open metal sites, ultra-microporous MOFs, and amine functionalized MOFs. The mechanisms combined with trace CO2 capture performance by these MOFs are discussed in detail, offering some promising adsorption solutions for future practical applications of MOF materials. In addition, the performance for CO2 capture under humid conditions and the regenerability of these MOF adsorbents are revealed. In order to address major issues of capacity, selectivity and stability especially under humid conditions, precise construction and engineering of MOFs to achieve the optimized porous materials are needed.
Article
Natural gas is an environmentally friendly energy resource and it is of central importance to remove acid gases (e.g. H2S and CO2) from natural gas. In this study, we computationally screen 6013 metal–organic frameworks (MOFs) for the simultaneous separation of H2S and CO2 from natural gas under humid conditions (represented by a six-component CH4/C2H6/C3H8/H2S/CO2/H2O gas mixture). To minimize the competitive adsorption of H2O, 606 hydrophobic MOFs are first selected on the basis of the Henry constants of H2O and subsequently assessed for the adsorption capacity of H2S + CO2 (NH2S+CO2) and the selectivity of H2S + CO2 over C1–C3 (SH2S+CO2/C1–C3). Structure–performance relationships are established between MOF descriptors (the largest cavity diameter, surface area, void fraction and isosteric heat) and performance metrics (NH2S+CO2 and SH2S+CO2/C1–C3). Moreover, a new performance metric (TSN, the trade-off between NH2S+CO2 and SH2S+CO2/C1–C3) is proposed. From the Pearson correlation technique, it is revealed that the TSN has strong correlations with the MOF descriptors. Finally, the best MOFs are identified. Interestingly, most of the best MOFs contain N-rich organic linkers (e.g. pyridine and azoles), which are recommended for the design of functional MOFs for H2S and CO2 separation. The microscopic insights and guidelines provided by this computational study are useful toward the development of new MOFs for the efficient upgrading of natural gas.