Next Article in Journal
Theoretical Study of the Magnetic Properties of the SmFe12−xMox (x = 1, 2) and SmFe10Mo2H Compounds
Previous Article in Journal
Dynamics of Core–Shell-Structured Sorbents for Enhanced Adsorptive Separation of Carbon Dioxide
Previous Article in Special Issue
Efficient Adsorption Removal of Tetrabromobisphenol A from Water by Using a Magnetic Composite Fe3O4/GO/ZIF-67
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exploring Methane Storage Capacities of M2(BDC)2(DABCO) Sorbents: A Multiscale Computational Study

by
Nguyen Thi Xuan Huynh
1,2,*,
Tue Nguyen-Van
3,
Nguyen Le Bao Tran
1,2,
Nguyen Van Nghia
1 and
Pham Ngoc Thanh
4
1
Department of Physics and Materials Science, Faculty of Natural Sciences, Quy Nhon University, 170 An Duong Vuong, Quy Nhon 55000, Vietnam
2
Lab of Computational Chemistry and Modelling (LCCM), Quy Nhon University, 170 An Duong Vuong, Quy Nhon 55000, Vietnam
3
ExploraScience Quy Nhon, 10 Science Avenue, Quy Nhon 55000, Vietnam
4
Department of Precision Engineering, Graduate School of Engineering, Osaka University, 2-1, Yamada-oka, Suita, Osaka 565-0871, Japan
*
Author to whom correspondence should be addressed.
Crystals 2024, 14(7), 596; https://doi.org/10.3390/cryst14070596
Submission received: 21 May 2024 / Revised: 24 June 2024 / Accepted: 25 June 2024 / Published: 27 June 2024

Abstract

:
A promising solution for efficient methane (CH4) storage and transport is a metal–organic framework (MOF)-based sorbent. Hence, searching for potential MOFs like M2(BDC)2(DABCO) to enhance the CH4 storage capacity in both gravimetric and volumetric uptakes is essential. Herein, we systematically elucidate the adsorption of CH4 in M2(BDC)2(DABCO) or M(DABCO) (M = Mg, Mn, Fe, Co, Ni, Cu, Zn) MOFs using multiscale simulations that combined grand canonical Monte Carlo simulation with van der Waals density functional (vdW-DF) calculation. We find that, in the M(DABCO) series, Mg(DABCO) has the highest total CH4 adsorption capacities, with m t o t = 231.39 mg/g at 298 K, for gravimetric uptake, and V t o t = 231.43 cc(STP)/cc, for volumetric uptake. The effects of temperature, pressure, and metal substitution on enhancing CH4 storage are evaluated, and we predict that the volumetric CH4 storage capacity on M(DABCO) could meet the DOE target at temperatures of ca. 238 K–268 K and pressures of 35–100 bar. The interactions between CH4 and M(DABCO) are dominated by the vdW interactions, as shown by the vdW-DF calculations. The Mg, Mn, Fe, Co, and Ni substitutions in M(DABCO) result in a stronger interaction and thus, a higher CH4 storage capacity, at higher pressures for Mg, Mn, Ni, and Co and at lower pressures for Fe. This work may provide guidance for the rational design of CH4 storage in M2(BDC)2(DABCO) MOFs.

1. Introduction

Recently, the use of methane (CH4) has received increasing attention as a potential fuel to resolve the global energy crisis and to mitigate greenhouse emissions [1,2,3]. This is because CH4 exhibits a lower CO2 emission and the highest H:C ratio. Moreover, CH4 is abundant, as it is the major component of natural gas (about 85%) and is found in the emissions of fossil fuel combustion, caves, and deep rock wells. Therefore, CH4 can provide an energy source for hundreds of years [4]. Although the energy from CH4 is not entirely green like hydrogen, promising technologies for converting methane into basic chemicals will enable it to become a potential bridge fuel and the basis of organic substances in the future. Furthermore, methane can be obtained from synthesizing CO2 and water, which contributes to reducing carbon emissions globally [4]. Nevertheless, the low energy density is a main drawback of employing CH4 as a fuel in light-duty vehicles. The volumetric energy density of methane is nearly 900 times lower than that of gasoline at standard temperature and pressure (STP: the pressure of 1 atm and the temperature of 273.15 K) [5]. Sorbent-based technology is one potential solution to resolve this drawback and enhance the CH4 storage efficiency [5]. For efficient storage and distribution, methane fuel technology needs to meet three objectives [2,6,7,8]: (i) a total gravimetric storage density of ( 500   m g C H 4 / g s o r b e n t or 700 cm3/g), (ii) a total container density of ( 155   m g / c m 3   o r   263 cc(STP)/cc), and (iii) a storage condition in the range of near ambient temperatures and moderate pressures (typical pressure is under 35 bar, but storage pressure is up to 65 bar). It is also important to note that when factoring in the 25% packaging loss, the capacity needs to reach 330 cc(STP)/cc for the container [6]. Therefore, finding suitable sorbents for methane storage is attracting numerous researchers.
Among sorbents, metal–organic frameworks (MOFs) have received significant attention for CH4 storage because they exhibit several prominent properties for CH4 adsorption, e.g., high thermal and physical stability, changeable porosity, adjustable organic functionality, structural flexibility, and high-pressure resistance, etc. [2,9,10,11]. MOFs have been of interest in energy storage, especially methane storage, because of their significant enhancement of the volumetric density of methane [12].
Many works reported the methane storage capacity of sorbents, including MOFs, at room temperature [5,12,13,14,15,16]. Among them, several impressive MOFs recorded the highest CH4 storage (gravimetric uptake or volumetric uptake) over time, such as Co2(BPY)3(NO3)4 (1997) [17], IRMOF-6 (2002) [18], HKUST-1 (2013) [6], Al-soc-MOF-1 (2015) [19], Cu-tbo-MOF-5 (2016) [20], and MFM-180 and MFM-185 (2017) [21]. Generally, at room temperature, the CH4 storage in MOFs with outstanding gravimetric CH4 uptake exhibits relatively low volumetric uptake or vice versa. So far, Al-soc-MOF-1 was reported to have the highest total gravimetric CH4 storage capacity, with 579 (362) cm3/g ca. 420 (263) mg/g at 65 (35) bar. However, the volumetric uptake of CH4 in Al-soc-MOF-1 was only 197 cc/cc at 65 bar (123 cc/cc at 35 bar) [19]. It is speculated that an ultra-large pore volume and a high surface area of Al-soc-MOF-1 are responsible for its outstanding gravimetric CH4 storage.
Shortening the organic linker in Al-soc-MOF-1 results in enhanced volumetric CH4 storage due to the strengthening of the CH4–MOF framework interaction. We noticed that the available data shows that gravimetric CH4 uptake in MOF, with high surface area and pore volume, is always high, but its volumetric uptake is rather low due to a weak CH4–MOF interaction (Table S1) [19,22,23,24]. In contrast, MOFs, having a moderate surface area and pore volume, show a high volumetric, but a limited gravimetric, uptake [8,25]. To date, only a few MOFs have simultaneously achieved high gravimetric and volumetric CH4 storage capacities at room temperatures and moderate pressures under 100 bar. Some of the best MOFs that meet these criteria are NJU-Bai 43 [26], UTSA-76 [27], and UTSA-110a [28]. Therefore, finding MOFs that achieve both high gravimetric and volumetric adsorption simultaneously, at moderate temperature and pressure, is a challenging task.
The M2(BDC)2(DABCO) [M(DABCO)] MOFs are promising sorbents for many applications in the fields of gas storage and capture, including outstanding CH4 storage capacity [29,30,31,32,33,34,35]. The M2(BDC)2(DABCO) consists of a metal node connected via the double ligands (BDC = benzene dicarboxylate and DABCO = 1,4-diazabicyclo [2,2,2] octane) (see Figure 1). BDC is among the most popular linkers for forming MOFs with simple cubic topologies [36]. Additionally, DABCO-containing materials are expected to exhibit unique physicochemical effects and properties [37]. So far, M(DABCO) with M = Zn, Co, Cu, Ni, and Fe has been successfully synthesized [38,39,40,41,42], in which M(DABCO) (M = Zn, Co, Cu, Ni) has shown high gravimetric and volumetric CH4 sorption capacities at room temperatures (see Table S2) [38,39,40,41]; however, CH4 adsorption in Fe(DABCO) has not yet been evaluated [42]. Although some experimental results have shown that M(DBACO) MOFs obtain high CH4 adsorption capacity, the roles of metal substitution in enhancing CH4 uptakes in both volumetric and gravimetric units remain elusive. Previous works show that metal substitution is an effective way to tune the CH4 capacity in MOFs [5,43], as MOFs contain paddlewheel units, such as M-HKUST-1, and M2(BDC)2(TED) enhances CH4 adsorption compared to many other MOFs, especially in regards to volumetric density [6,8,41,44]. Among paddle-wheel MOFs, the M-HKUST-1 series (M = transition metals or alkaline earth metals) has been experimentally studied and theoretically predicted to exhibit potential for various applications, including gas adsorption [45,46,47,48,49]. More importantly, several available studies show that M-modified HKUST-1 (M = Mg, Mn) or porph@MOM MOFs (M = Mg, Mn, Fe, Co, Ni), having the same space group and tbo topology as HKUST-1 with paddlewheel units, have been successfully synthesized, enhancing the efficiency of gas adsorption or gas reduction [50,51,52,53]. Meanwhile, research on metal-substituted M(DABCO) used to enhance CH4 storage capacity has received less attention than has HKUST-1. We also note that experimentally, the CH4 adsorption on available M(DABCO) is studied in a limited temperature and pressure range.
For the above reasons, we selected the metals, including Co, Ni, Cu, Zn, and Fe, achieving the experimentally synthesized M(DABCO) and other potential predicted metals (Mg, Mn) in M(DABCO) to explore the methane adsorption capacity. Herein, we employed multiscale simulation, combining grand canonical Monte Carlo simulation and density functional theory calculations to study the effect of metal substitution on CH4 uptake in M(DABCO) and to gain physical insights into the interactions between CH4 and M(DABCO).

2. Methods and Modeling

2.1. Modeling of M2(BDC)2(DABCO) MOFs

The configuration of M2(BDC)2(DABCO) or M2(BDC)2(TED) is shown in Figure 1. Hereafter, we denote M2(BDC)2(DABCO) as M(DABCO) for simplicity. The chemical formula of M(DABCO) is C22H20N2O8M2. The unit cell of M(DABCO) consists of divalent metal ions M2+ (M = Mg, Mn, Fe, Co, Ni, Cu, Zn) and two linkers, namely DABCO and BDC. DABCO is 1,4-diazabicyclo[2,2,2]octane with the formula of C6H12N2, exhibiting a cage-like structure; BDC2− is 1,4-benzenedicarboxylate, with the formula of C8H4O4. The lattice constants and atomic positions of M(DABCO) were fully optimized using the DFT calculation (the DFT detail will be provided later) and the Murnaghan equation of state [54,55,56].

2.2. Grand Canonical Monte Carlo Simulation

The adsorbed amount of CH4 (methane) and the adsorption isotherms of N2 (nitrogen) and CH4 in M(DABCO) were computed using grand canonical Monte Carlo (GCMC) simulation at room temperature and various pressures up to 100 bar using the RASPA code [57]. The GCMC simulations were performed using the μVT ensemble (constant chemical potential, volume, and temperature). First, we performed 1.5 × 105 equilibration cycles and then performed 3.5 × 105 MC steps for random insertion, random deletion, translation, or rotation of molecules (N2, CH4) in the M2(BDC)2(DABCO) box (repeated three times, with the primitive unit cell in three dimensions) to obtain the equilibrium systems.
The van der Waals interaction mainly dominates the interaction between CH4 and the MOF, while electrostatic interaction plays a minor role [3]. Therefore, the interaction between CH4 and M(DABCO) is described by the 12-6 Lenard Jones (LJ) potential for the van der Waals interaction. We further verified this by elucidating the point charge of CH4 adsorption in M(DABCO) using the DFT calculation. The charge of methane is thus modeled as an uncharged single-atom model. Therefore, the interaction between the atoms of the CH4 molecule and the M(DABCO) is calculated as
V = ( i , j ) V i j = ( i , j ) 4 ε i j σ i j r i j 12 σ i j r i j 6 ,
where r i j denotes the distance between the atoms i and j . The parameters σ i j and ε i j were determined by the mixing rule of Lorentz and Berthelot [58,59] as
σ i j = 1 2 σ i + σ j ,   ε i j = ε i ε j 1 / 2 .
Herein, the parameters σ i and ε i , the size and strength parameters, are selected from the generic force fields, which are tabulated in Table 1 [57]. The V i j was truncated with the cutoff radius ( r c u t ) of 20 Å. During GCMC simulations, the M(DABCO) framework was fixed, while CH4 could move freely in the pores of the MOF until reaching the equilibrium state. We modeled the CH4 molecules using the TraPPE force field (also known as TraPPE-UA), developed by Martin et al. [60]. The accuracy of our approach is validated because it shows a good agreement between the GCMC and the experimental results for M(DABCO) (M = Co, Cu, Ni, Zn) [38,39,41] (Figure S1). Furthermore, a good agreement between the GCMC and the experimental results has been shown for previous flexible MOFs such as Co3(NDC)3(DABCO) and MIL-53 [3,61]. Here, it should also be noted that the results obtained via adsorption equilibrium using GCMC simulations are predictive.
For evaluating the adsorption capacity of CH4 on M(DABCO) (M = Mg, Mn, Fe, Co, Ni, Cu, Zn), GCMC simulations were performed with various pressures up to 100 bar and at room temperature. Here, we determined both gravimetric and volumetric uptakes of CH4 in M(DABCO). We denoted total and excess gravimetric uptakes as m t o t and m e x c , respectively. On the other hand, total and excess volumetric uptakes are denoted as V t o t and V e x c , respectively. The total uptake ( m t o t ) and excess uptake ( m e x c ) are related through the equation:
m t o t = m e x c + d C H 4 × V p
where, d C H 4 is methane density, and V p is the total pore volume of M(DABCO). In order to elucidate the strength of the CH4-M(DABCO) interaction, the total heat of adsorption was calculated using the following expression [62]
Q s t = H a d s = U R T ,
Herein, U is the internal energy, which is estimated by the formula:
U = U h g U h U g
where U h , U g , and U h g are the average energy of the M(DABCO) (host), the CH4 (gas), and the (host + gas) system, respectively. T is the temperature, and R is the gas constant.
The specific surface area ( A B E T ) and pore volume ( V p ) are important structural factors of M(DABCO). The A B E T and V p of M(DABCO) were determined from the nitrogen (N2) adsorption isotherms at 77 K, which can be considered similar to the experimental sample using the BET (Brunauer–Emmett–Teller) model. The pore-size distribution (PSD) was calculated geometrically using Gelb and Gubbins’ method [57].

2.3. Density Functional Theory Calculation

We employed the DFT calculation to optimize the M(DABCO) structures and to accurately elucidate the interaction between CH4 and M(DABCO) [64,65,66]. These DFT calculations were performed using the Vienna ab initio simulation package (VASP) code [67,68], with the projected augmented wave (PAW) method [69,70] and the plane-wave basis set, with a cut of energy of 700 eV. The van der Waal density functional (vdW-DF) method, with the revised Perdew–Burke–Ernzerhof exchange-correlation functionals [71], was employed to accurately describe the CH4-M(DABCO) interaction, denoted as CH4@M(DABCO). The Brillouin zone was sampled using Monkhorst–Pack meshes [72] of 3 × 3 × 3, for atomic optimization, and 5 × 5 × 5, for calculating the total energies and electronic properties (charges, density of state, etc.). The atomic charge was elucidated using charge Bader partitioning [73,74], while electronic structure analysis was performed using the density of states (DOS). The adsorption energy (Eads) of the CH4 molecule adsorbed on the M(DABCO) was estimated as
E a d s = E M O F + C H 4 ( E M O F + E C H 4 )
Here, E M O F + C H 4 , E M O F , and E C H 4 are the total energies of the CH4 adsorbed-MOF, the pristine M(DABCO), and the isolated methane molecule, respectively.

3. Results and Discussion

3.1. Porosity and Surface Analysis of M(DABCO)

Firstly, we discuss the porosity and surface area of the M(DABCO) frameworks. The N2 sorption isotherms in M(DABCO) and their pore-size distribution (PSD) are shown in Figure 2a and Figure 2b, respectively. These N2 sorption isotherms indicate the type-I adsorption isotherms, similar to the N2 isotherm on Zn(BDC)(DABCO)0.5 [75] and Co(BDC)(DABCO)0.5 [39] (these increased rapidly at a pressure near 0 bar and saturated at ca. 500 cm3/g). Herein, the results show that Mg(DABCO) is superior to the remaining M(DABCO) MOFs in regards to N2 adsorption.
The results of the geometrical features of M(DABCO) are listed in Table 2. Our results for PSD, pore size, surface area, and pore volume are relatively consistent with the experimental reference values for M(DABCO) (M = Zn, Co) [39,75]. The surface area and pore volume are in ascending order of metal: Cu < Fe < Zn < Co < Ni < Mn < Mg, in which Cu (1561 m2/g, 0.71 cm3/g) and Mg (1931 m2/g, 0.87 cm3/g) exhibit the smallest and the largest surface area and pore volume of M(DABCO), respectively. The results are comparable with the experimental data [30,39,75,76], where the A B E T and V p of Ni exhibited the largest values in the experimentally measured metals (Cu, Co, Zn, and Ni) (Table 2). From the data for A B E T and V p in Table 2, we also find the linear relationship between A B E T and V p , i.e., A B E T = a · V p , with the fitted linear coefficient ( a ) equal to 2204 ± 6 , with an error of 0.25% (Figure S2). Moreover, we also calculate the pore sizes of M(DABCO), and they are about 8.45 to 8.72 Å, which are close to the results in the experiment data (ca. 8.0 Å) (see Table 2).

3.2. Methane Adsorption on M(DABCO)

Because the force field parameters are important to accurately evaluate the adsorption capacity of CH4 on M(DABCO) (M = Mg, Mn, Fe, Co, Ni, Cu, Zn) in GCMC simulation, we first validate our approach using available experimental CH4 uptake results for Co(DABCO), Cu(DABCO), Ni(DABCO), and Zn(DABCO) [38,39,41] (Figure S1). The results indicate that our GCMC results using the TraPPE-UA force field for CH4 and the LJ parameters for MOFs, taken from the universal force field (UFF) for metals and the DREIDING force field for other atoms, achieve reliability compared to the experimental data [38,39,41]. In particular, under the same conditions, with a temperature of 303 K and a pressure of 75 bar, the simulation indicates that the CH4 uptake of Co(DABCO) is m e x c = 139 mg/g, compared to that of Co(DABCO) measured by the experiment (140 mg/g), with an error of ca. 0.7% [39]. Additionally, under the same conditions of 298 K and 35 bar, the CH4 adsorption of Cu(DABCO) achieved 181 cm3/g by our simulation and 186 cm3/g by the experiment in Ref. [38], with an error of ca. 3.2%. Moreover, our GCMC simulations are more consistent with the experimental results than those carried out by the force field proposed by Dubbeldam et al. [62,63].
The CH4 storage capacities are shown in Figure 3, for low pressures under 1 bar, and in Figure 4, for higher pressures up to 100 bar, respectively. Among these MOFs, under 1 bar, Mg(DABCO) yields the largest gravimetric uptakes, with m t o t = 15.15 mg/g and m e x c = 14.58 mg/g. Moreover, Fe(DABCO) expresses the largest volumetric uptakes w i t h   V t o t = 16.98 cc/cc and V e x c = 16.42 cc/cc. On the other hand, Ni(DABCO) exhibits the lowest CH4 uptakes. However, the deviation between the highest and lowest M(DABCO) for the gravimetric [volumetric] CH4 uptake is small, with a difference of 2.5 mg/g (2.6 cc/cc) when comparing the Ni(DABCO) results with those for Mg(DABCO) [Fe(DABCO)].
When the pressure increases to 100 bar, the difference in the CH4 adsorption capacity of M(DABCO) becomes more apparent (Figure 4). The total CH4 sorption isotherms show a strong dependence on pressure up to 100 bar, while excess CH4 sorption isotherms nearly saturate at 35 bar, then increase slightly and reach maximum values at about 60–65 bar. Methane storage capacities are listed in detail in Table 3. For comparison with available experimental data, we listed the most outstanding adsorption capacities of CH4 on MOFs for the gravimetric and volumetric uptakes in Table 3. Our results for the CH4 storage in M(DABCO) show that the gravimetric uptakes of M(DABCO) [182−231 mg/g] are lower than that of Al-soc-MOF-1 [420 mg/g], recorded the highest gravimetric uptake thus far. However, comparing to the volumetric uptake of Al-soc-MOF-1 (197 cc/cc), the CH4 volumetric uptakes of M(DABCO) [220–231 cc/cc] are larger, showing that M(DABCO) is a promising candidate for CH4 storage in terms of the volumetric uptake. This can be further confirmed by comparing both uptakes of the M(DABCO) series and HKUST-1, which is recorded to yield the highest volumetric capacity, along with a relatively high gravimetric uptake, in the literature. The CH4 storage in HKUST-1 is 216 mg/g and 267 cc/cc at 65 bar, which is comparable with that noted in the M(DABCO) series (182–231 mg/g and 220–231 cc/cc) at 100 bar. Compared to the best available data (Table 3), our results show that the CH4 adsorption capacities on M(DABCO) MOFs are remarkably enhanced, especially in regards to their volumetric density.
Here, we also discuss the effect of metal substitution on CH4 storage capacities in M(DABCO). The CH4 adsorption capacity of M(DABCO) varies in the order of Cu(DABCO) < Fe(DABCO) < Zn(DABCO) < Co(DABCO) < Ni(DABCO) < Mn(DABCO) < Mg(DABCO). Among them, Cu(DABCO) yields the lowest CH4 adsorption capacity in regards to both mass and volume density, while Mg(DABCO) achieves the highest gravimetric and volumetric CH4 adsorption capacity. The results for Mg(DABCO) are m t o t = 231.39 mg/g and m e x c = 172.30 mg/g; V t o t = 231.43 cc/cc and V e x c = 172.33 cc/cc in STP. The volumetric uptake of Mg(DABCO) is higher than that of Al-soc-MOF-1 and is comparable to that of HKUST-1. To identify the best gravimetric CH4 adsorption of Mg(DABCO) compared to other M(DABCO), we evaluated the total amount of CH4 adsorption per mole of M(DABCO) at 298 K and 100 bar and the molar mass of M(DABCO). The results are shown in Figure S3 and Table S3. The highest gravimetric CH4 uptake in Mg(DABCO) arises from both a smaller molar mass and a higher amount of CH4 adsorption per mole of Mg(DABCO). We find that the molar mass effect is dominant when comparing the results of Mg(DABCO) with Mn(DABCO) and Ni(DABCO), showing that light metal substitution is an effective strategy to enhance gravimetric CH4 uptake.
Although the results obtained for adsorption capacities, especially volumetric uptakes, of CH4 in M(DABCO) are remarkable, they have not yet reached the DOE’s standards (500 mg/g and 263 cc/cc). Therefore, we further elucidate CH4 uptakes in M(DABCO) in the temperature range from 298 K to 238 K to strengthen CH4 adsorption. Here, we only considered Mg(DABCO), which is the best candidate for CH4 storage in the M(DABCO) series. The results are shown in Figure 5a. If the packing efficiency loss of CH4 is ignored, at 238 K and 35 bar, V t o t   i s   270.61 cc/cc, which could meet the DOE target (263 cc/cc). At higher pressures (65 bar and 100 bar), the temperatures at which the V t o t reaches the DOE standard are 258 K and 268 K, respectively, showing that elevating the pressure could increase the temperature at which the volumetric uptake meets the DOE requirement. We also note that this relationship between pressure and temperature is linear (see details in Table 4). Figure 5b shows that the total CH4 uptakes (at 35 bar, 65 bar, and 100 bar) and the maximum excess CH4 uptake depend linearly on the temperature. These trends could be seen in many MOFs for methane uptakes available at 1 bar, especially for CH4 adsorbed on M(BDP) (M = Co, Fe; BDP = 1,4-benzenedipyrazolate) [13].
We further find a correlation between the CH4 uptakes with the porosity and surface area of M(DABCO). The M(DABCO), with a higher surface area, shows a higher gravimetric uptake (Figure 6a), while M(DABCO), with a higher pore volume, exhibits a higher volumetric uptake (Figure 6b). Figure 6a shows the total and excess gravimetric uptake of CH4 adsorbed on M(DABCO), as a function of its surface area ( A B E T ). The results show a linear behavior between gravimetric uptakes and surface areas. Similarly, we also find a linear relationship between the volumetric uptakes of CH4 in M(DABCO) and its pore volume. These relationship trends are also found in the high-pressure range in the literature [8,19]. This shows that improving the structural features ( A B E T and V p ) of the MOF increases the storage ability of CH4 on M(DABCO), especially in regards to gravimetric uptakes.

3.3. Adsorption Heat of Methane on M(DABCO)

In order to elucidate the average interaction between the CH4 molecules and the M(DABCO) MOFs (CH4@M(DABCO)), the isosteric heat of adsorption ( Q s t = H a d s ) at room temperature was calculated and visualized (see Figure 7). The results reveal that the values of Q s t for CH4@M(DABCO) are about 16 to 18 kJ/mol, with the average of Q s t ≈ 17 kJ/mol. Generally, the heat of adsorption of CH4 physisorption is about 15 to 25 kJ/mol for sorbents recording the highest volumetric CH4 capacities [77], and the Q s t of CH4@M(DABCO) is also within this range. Hence, we conclude that M(DABCO) is promising for CH4 storage, based on physisorption. It is also noted that at the pressures below 1 bar, the Q s t of CH4@Fe(DABCO) is also higher than that of the remaining metals, explaining the stronger CH4 adsorption on Fe(DABCO) sorbents below 1 bar (the inset of Figure 7). The relatively small difference between the adsorption heats of M(DABCO) explains the reason for the insignificant change in the volumetric CH4 adsorption capacity of M(DABCO) (Figure 4b).

3.4. Stable CH4 Adsorption Sites on M(DABCO)

To gain insight into the stable adsorption sites of CH4@M(DABCO), we search the adsorption sites of CH4 in M(DABCO) and evaluate the adsorption energies of the CH4@M(DABCO) based on vdW-DF calculations. Here, we aim to reveal the influence of the metal substitution in M(DABCO) in regards to CH4-M(DABCO) interaction. Hence, we considered two adsorption sites of CH4, i.e., site 1, where CH4 is adsorbed at the metal oxide cluster site (Figure S4), and site 2, where CH4 is located at the interface region between the M-O-C cluster and the TED group (Figure S5).
The adsorption energies ( E a d s ) and the nearest distance between CH4 and M(DABCO) of the CH4@M(DABCO) adsorption configurations are tabulated in Table 5. The average interaction strength of CH4@M(DABCO) for two adsorption sites is in order of Mn > Co ≈ Mg ≈ Fe > Ni > Zn > Cu. The E a d s of CH4@M(DABCO), with M = Mn, Co, Mg, Fe, and Ni, are similar (–73 to –78 kJ/mol), whereas the E a d s of CH4@Zn(DABCO) are slightly smaller (–62 kJ/mol). In contrast, the E a d s of CH4@Cu(DABCO) are the smallest (–20.31 and –45.33 kJ/mol for sites 1 and 2, respectively). The nearest distances (d1 and d2) between CH4 and M(DABCO) are more than 2.5 Å (Table 5). Our calculated Eads, d1, and d2 of CH4@M(DABCO) show that the interactions between CH4 and the M(DABCO) series are primarily physisorption interactions, in agreement with the heat of adsorption recorded in the GCMC simulations. We find that the CH4@M(DABCO) interactions are almost unaltered by the metal substitutions because this change in E a d s is small for Mn, Co, Mg, Fe, Ni, and Zn. In contrast, Cu yields a weaker CH4 binding to Cu(DABCO) [78,79].
Bader charges for the adsorption sites are listed in Table 6 and Table 7, where symbols C1, C2, and C3 represent the C atoms linked with other C, O, and N atoms of MOF, respectively (Figure S6). The results show a tiny charge exchange between the atoms of CH4 and M(DABCO), within the error range of 0.005 e in the Bader analysis. The magnitude of the Bader charge of C in CH4 is less than 0.038 e, whereas that of four H atoms in CH4 is less than 0.037 e. Overall, the CH4 adsorbed on CH4@M(DABCO) is almost neutral because the total Bader charge of CH4 is almost zero. Therefore, the contribution of the electrostatic interaction between CH4 and M(DABCO) could be ignored. This result further validates our GCMC approach, in which the interaction between CH4-M(DABCO) is modeled by LJ potentials, leading to a good agreement between our GCMC results and those of the available experimental data from Refs. [38,39,41] (Figure S1).
A DOS analysis of CH4 and M(DABCO) shows that the 1 t 2 (or σ p ) state of CH4 is the main contributor to the interaction between CH4 and M(DABCCO). However, these changes in DOS are not significant, and include mainly the fact that the DOS peaks are shifted toward the negative energy, in which the 1 α 1 (or σ s ) state barely changes its shapes after CH4 adsorption (Figure 8). The slight difference in adsorption energies can also explain the small change in volumetric CH4 adsorption capacity when substituting metal in M(DABCO) due to the good correlation between the CH4 uptake and the adsorption energy on MOFs [5]. The origin of the strongest and weakest interactions of CH4 in Mn(DABCO) and Cu(DABCO) can be seen from the overlap between the DOS of CH4 and the DOS of M(DABCO) (Figures S7 and S8). Generally speaking, a larger overlap in the DOS of CH4 and the MOF framework gives rise to a stronger interaction [54]. We see that the overlap between the DOS of CH4 and Cu(DABCO) is the smallest, while that of CH4 and Mn(DABCO) is the largest, confirming a difference in the binding strength of CH4 to these frameworks.
Finally, we discuss the stability of M(DABCO) with M = Mg and Mn because these MOFs have not yet been synthesized by experiments. To this end, we elucidated the enthalpies of the formation of M(DABCO) with M = Mg and Mn, and the results are shown in Table S4. We find that the enthalpies of formation of M(DABCO) with M = Mg and Mn are negative, showing that at least these MOFs are stable with respect to isolated metal ions and their constituent organic linkers. Therefore, our calculated results raise the possibility of those MOF formations, and we encourage experiments to explore the synthesis of M(DABCO) with M = Mg and Mn due to their remarkable CH4 storage capacity and stable CH4 adsorption. Recently, Mn-HKUST-1 or Mn- and Mg-based MOFs like HKUST-1 with similar paddlewheel building units have been successfully synthesized, further supporting our hypothesis [50,53].

4. Conclusions

Using a multiscale simulation that combined GCMC and DFT, this work elucidated the CH4 storage capacity of the M(DABCO) series (M = Mg, Mn, Fe, Co, Ni, Cu, Zn). We obtained the following important results:
(i) 
The CH4 storage capacities in regards to both the gravimetric and volumetric uptakes of the M(DABCO) series are evaluated, and the volumetric uptakes of CH4 in M(DABCO) are remarkable. Both uptakes are in the increasing order of M(DABCO), as: Cu < Fe < Zn < Co < Ni < Mn < Mg. Among these MOFs, Mg (DABCO) shows the highest CH4 adsorption, with the maximum total and excess uptakes of 231.39 mg/g (at 100 bar) and 172.30 mg/g (at 65 bar), for gravimetric uptakes, and 231.43 cc(STP)/cc (at 100 bar) and 172.33 cc(STP)/cc (at 65 bar), for volumetric uptakes. Moreover, the methane adsorption capacities increase almost linearly with temperature, and the gravimetric methane adsorption strongly depends on the geometrical features (specific surface area and pore volume) of the M(DABCO) MOFs. Our simulations predicted that the volumetric CH4 storage capacity could meet the DOE target at lower temperatures, ca. 238 K–268 K in the 35–100 bar pressure range.
(ii) 
The interaction between CH4 and M(DABCO) is mainly governed by the vdW interaction, and electrostatic interactions play a minor role, as indicated by the DFT calculations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/cryst14070596/s1, Table S1: Several well-known MOFs for gravimetric CH4 adsorption capacity at 298 K; Table S2: Excess gravimetric and volumetric CH4 uptakes adsorbed on M(DABCO) at room temperature; Figure S1: Isotherms for methane compared with available experimental (exp.) data; Figure S2: The correlation between the surface area ( A B E T ) and the pore volume ( V p ) of M(DABCO) sorbents. Points indicate the pairs of V p and A B E T . The solid line indicates the linear fitting; Figure S3: The simulated CH4 adsorption isotherms of M(DABCO) with the total uptake (solid lines) and the excess uptake (dashed lines) at 298 K and pressures up to 100 bar (in CH4 molecules/unit cell); Table S3: The amount of CH4 adsorption in M(DABCO) at 100 bar, 298 K in many different units compared to the molar mass of M(DABCO); Table S4: Formation enthalpy ( H f ) of M2(BDC)2(TED) or M(DABCO) compounds; Figure S4: The most stable CH4 adsorption sites on the metal cluster of M(DABCO); Figure S5: The most stable CH4 adsorption sites on the interface between the M-O-C cluster and TED of M(DABCO); Figure S6: The symbols for the atoms (C1, C2, C3, O, N, and M) of M(DABCO) with M = Mg, Mn, Fe, Co, Ni, Cu, or Zn. Here, H atoms are omitted; Figure S7: The overlap between the DOS of CH4 and the atoms of M(DABCO) on the adsorption site of the metal cluster; Figure S8: The overlap between the DOS of CH4 and the atoms of M(DABCO) on the adsorption site of the M-O-C cluster—TED interface.

Author Contributions

Conceptualization, N.T.X.H. and P.N.T.; methodology, N.T.X.H. and T.N.-V.; validation, T.N.-V., N.L.B.T., N.V.N. and P.N.T.; formal analysis, T.N.-V., N.L.B.T. and N.V.N.; investigation, N.T.X.H., T.N.-V. and N.L.B.T.; resources, N.T.X.H., T.N.-V. and P.N.T.; data curation, T.N.-V. and N.L.B.T.; writing—original draft preparation, N.T.X.H., T.N.-V. and N.L.B.T.; writing—review and editing, N.T.X.H., N.V.N. and P.N.T.; visualization, T.N.-V., N.L.B.T. and N.V.N.; supervision, N.T.X.H. and P.N.T.; project administration, N.T.X.H.; funding acquisition, N.T.X.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research is funded by Vietnam Ministry of Education and Training (MOET), grant number B2022-DQN-05.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding author.

Acknowledgments

We acknowledge support for the calculations from the Lab of Computational Chemistry and Modeling (LCCM), Quy Nhon University, Quy Nhon, Vietnam.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. He, Y.; Zhou, W.; Qian, G.; Chen, B. Methane Storage in Metal-Organic Frameworks. Chem. Soc. Rev. 2014, 43, 5657. [Google Scholar] [CrossRef] [PubMed]
  2. Huong, T.T.T.; Thanh, P.N.; Huynh, N.T.X.; Son, D.N. Metal—Organic Frameworks: State-of-the-Art Material for Gas Capture and Storage. VNU J. Sci. Math.—Phys. 2016, 32, 67–85. [Google Scholar]
  3. Ribeiro, R.P.P.L.; Mota, J.P.B. Surface Area and Porosity of Co3(ndc)3(dabco) Metal-Organic Framework and Its Methane Storage Capacity: A Combined Experimental and Simulation Study. J. Phys. Chem. C 2021, 125, 2411–2423. [Google Scholar] [CrossRef]
  4. Arutyunov, V.; Savchenko, V.; Sedov, I.; Arutyunov, A.; Nikitin, A. The Fuel of Our Future: Hydrogen or Methane? Methane 2022, 1, 96–106. [Google Scholar] [CrossRef]
  5. Koh, H.S.; Rana, M.K.; Wong-Foy, A.G.; Siegel, D.J. Predicting Methane Storage in Open-Metal-Site Metal-Organic Frameworks. J. Phys. Chem. C 2015, 119, 13451–13458. [Google Scholar] [CrossRef]
  6. Peng, Y.; Krungleviciute, V.; Eryazici, I.; Hupp, J.T.; Farha, O.K.; Yildirim, T. Methane Storage in Metal-Organic Frameworks: Current Records, Surprise Findings, and Challenges. J. Am. Chem. Soc. 2013, 135, 11887–11894. [Google Scholar] [CrossRef] [PubMed]
  7. Granja-DelRío, A.; Cabria, I. Exploring the Hydrogen and Methane Storage Capacities of Novel DUT MOFs at Room Temperature: A Grand Canonical Monte Carlo Simulation Study. Int. J. Hydrogen Energy 2024, 54, 665–677. [Google Scholar] [CrossRef]
  8. Li, H.; Li, L.; Lin, R.-B.B.; Zhou, W.; Zhang, Z.; Xiang, S.; Chen, B. Porous Metal-Organic Frameworks for Gas Storage and Separation: Status and Challenges. EnergyChem 2019, 1, 100006. [Google Scholar] [CrossRef] [PubMed]
  9. Safaei, M.; Foroughi, M.M.; Ebrahimpoor, N.; Jahani, S.; Omidi, A.; Khatami, M. A Review on Metal-Organic Frameworks: Synthesis and Applications. TrAC—Trends Anal. Chem. 2019, 118, 401–425. [Google Scholar] [CrossRef]
  10. Li, D.; Yadav, A.; Zhou, H.; Roy, K.; Thanasekaran, P.; Lee, C. Advances and Applications of Metal-Organic Frameworks (MOFs) in Emerging Technologies: A Comprehensive Review. Glob. Chall. 2024, 8, 2300244. [Google Scholar] [CrossRef]
  11. Dutt, S.; Kumar, A.; Singh, S. Synthesis of Metal Organic Frameworks (MOFs) and Their Derived Materials for Energy Storage Applications. Clean Technol. 2023, 5, 140–166. [Google Scholar] [CrossRef]
  12. Sandhu, Z.A.; Raza, M.A.; Awwad, N.S.; Ibrahium, H.A.; Farwa, U.; Ashraf, S.; Dildar, A.; Fatima, E.; Ashraf, S.; Ali, F. Metal-Organic Frameworks for next-Generation Energy Storage Devices; a Systematic Review. Mater. Adv. 2024, 5, 30–50. [Google Scholar] [CrossRef]
  13. Ursueguía, D.; Díaz, E.; Ordóñez, S. Metal-Organic Frameworks (MOFs) as Methane Adsorbents: From Storage to Diluted Coal Mining Streams Concentration. Sci. Total Environ. 2021, 790, 148211. [Google Scholar] [CrossRef] [PubMed]
  14. Getman, R.; Bae, Y.; Wilmer, C.; Snurr, R. Review and Analysis of Molecular Simulations of Methane, Hydrogen, and Acetylene Storage in Metal–Organic Frameworks. Chem. Rev. 2012, 112, 703–723. [Google Scholar] [CrossRef] [PubMed]
  15. Guo, K. Applications of Metal-Organic Frameworks in Methane Storage. Highlights Sci. Eng. Technol. 2023, 58, 338–343. [Google Scholar] [CrossRef]
  16. Forrest, K.A.; Verma, G.; Ye, Y.; Ren, J.; Ma, S.; Pham, T.; Space, B. Methane Storage in Flexible and Dynamical Metal–Organic Frameworks. Chem. Phys. Rev. 2022, 3, 021308. [Google Scholar] [CrossRef]
  17. Kondo, M.; Yoshitomi, T.; Seki, K.; Matsuzaka, H.; Kitagawa, S. Three-Dimensional Framework with Channeling Cavities for Small Molecules: {[M2(4, 4′-Bpy)3(NO3)4]·xH2O}n (M=Co, Ni, Zn). Angew. Chemie Int. Ed. English 1997, 36, 1725–1727. [Google Scholar] [CrossRef]
  18. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O.M. Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science 2002, 295, 469–472. [Google Scholar] [CrossRef] [PubMed]
  19. Alezi, D.; Belmabkhout, Y.; Suyetin, M.; Bhatt, P.M.; Weseliński, L.J.; Solovyeva, V.; Adil, K.; Spanopoulos, I.; Trikalitis, P.N.; Emwas, A.H.; et al. MOF Crystal Chemistry Paving the Way to Gas Storage Needs: Aluminum-Based Soc -MOF for CH4, O2, and CO2 Storage. J. Am. Chem. Soc. 2015, 137, 13308–13318. [Google Scholar] [CrossRef]
  20. Spanopoulos, I.; Tsangarakis, C.; Klontzas, E.; Tylianakis, E.; Froudakis, G.; Adil, K.; Belmabkhout, Y.; Eddaoudi, M.; Trikalitis, P.N. Reticular Synthesis of HKUST-like Tbo-MOFs with Enhanced CH4 Storage. J. Am. Chem. Soc. 2016, 138, 1568–1574. [Google Scholar] [CrossRef]
  21. Moreau, F.; Kolokolov, D.I.; Stepanov, A.G.; Easun, T.L.; Dailly, A.; Lewis, W.; Blake, A.J.; Nowell, H.; Lennox, M.J.; Besley, E.; et al. Tailoring Porosity and Rotational Dynamics in a Series of Octacarboxylate Metal-Organic Frameworks. Proc. Natl. Acad. Sci. USA 2017, 114, 3056–3061. [Google Scholar] [CrossRef]
  22. Mason, J.A.; Veenstra, M.; Long, J.R. Evaluating Metal-Organic Frameworks for Natural Gas Storage. Chem. Sci. 2014, 5, 32–51. [Google Scholar] [CrossRef]
  23. Stoeck, U.; Krause, S.; Bon, V.; Senkovska, I.; Kaskel, S. A Highly Porous Metal–Organic Framework, Constructed from a Cuboctahedral Super-Molecular Building Block, with Exceptionally High Methane Uptake. Chem. Commun. 2012, 48, 10841–10843. [Google Scholar] [CrossRef] [PubMed]
  24. Li, B.; Zhang, Z.; Li, Y.; Yao, K.; Zhu, Y.; Deng, Z.; Yang, F.; Zhou, X.; Li, G.; Wu, H.; et al. Enhanced Binding Affinity, Remarkable Selectivity, and High Capacity of CO2 by Dual Functionalization of a Rht-Type Metal–Organic Framework. Angew. Chemie 2012, 124, 1441–1444. [Google Scholar] [CrossRef]
  25. Zhang, X.; Lin, R.B.; Alothman, Z.A.; Alduhaish, O.; Yildirim, T.; Zhou, W.; Li, J.R.; Chen, B. Promotion of Methane Storage Capacity with Metal-Organic Frameworks of High Porosity. Inorg. Chem. Front. 2022, 10, 454–459. [Google Scholar] [CrossRef]
  26. Zhang, M.; Zhou, W.; Pham, T.; Forrest, K.A.; Liu, W.; He, Y.; Wu, H.; Yildirim, T.; Chen, B.; Space, B.; et al. Fine Tuning of MOF-505 Analogues To Reduce Low-Pressure Methane Uptake and Enhance Methane Working Capacity. Angew. Chemie—Int. Ed. 2017, 56, 11426–11430. [Google Scholar] [CrossRef] [PubMed]
  27. Li, B.; Wen, H.M.; Wang, H.; Wu, H.; Tyagi, M.; Yildirim, T.; Zhou, W.; Chen, B. A Porous Metal-Organic Framework with Dynamic Pyrimidine Groups Exhibiting Record High Methane Storage Working Capacity. J. Am. Chem. Soc. 2014, 136, 6207–6210. [Google Scholar] [CrossRef] [PubMed]
  28. Wen, H.M.; Li, B.; Li, L.; Lin, R.B.; Zhou, W.; Qian, G.; Chen, B. A Metal–Organic Framework with Optimized Porosity and Functional Sites for High Gravimetric and Volumetric Methane Storage Working Capacities. Adv. Mater. 2018, 30, 1704792. [Google Scholar] [CrossRef] [PubMed]
  29. Liang, W.; Xu, F.; Zhou, X.; Xiao, J.; Xia, Q.; Li, Y.; Li, Z. Ethane Selective Adsorbent Ni(bdc)(ted)0.5 with High Uptake and Its Significance in Adsorption Separation of Ethane and Ethylene. Chem. Eng. Sci. 2016, 148, 275–281. [Google Scholar] [CrossRef]
  30. Xiang, H.; Ameen, A.; Gorgojo, P.; Siperstein, F.R.; Holmes, S.M.; Fan, X. Selective Adsorption of Ethane over Ethylene on M(bdc)(ted)0.5 (M = Co, Cu, Ni, Zn) Metal-Organic Frameworks (MOFs). Microporous Mesoporous Mater. 2020, 292, 109724. [Google Scholar] [CrossRef]
  31. Peng, L.; Wu, S.; Yang, X.; Hu, J.; Fu, X.; Huo, Q.; Guan, J. Application of Metal Organic Frameworks M(bdc)(ted)0.5 (M = Co, Zn, Ni, Cu) in the Oxidation of Benzyl Alcohol. RSC Adv. 2016, 6, 72433–72438. [Google Scholar] [CrossRef]
  32. Tan, K.; Canepa, P.; Gong, Q.; Liu, J.; Johnson, D.H.; Dyevoich, A.; Thallapally, P.K.; Thonhauser, T.; Li, J.; Chabal, Y.J. Mechanism of Preferential Adsorption of SO2 into Two Microporous Paddle Wheel Frameworks M(bdc)(ted)0.5. Chem. Mater. 2013, 25, 4653–4662. [Google Scholar] [CrossRef]
  33. Son, D.N.; Thuy Huong, T.T.; Chihaia, V. Simultaneous Adsorption of SO2 and CO2 in an Ni(bdc)(ted)0.5 Metal-Organic Framework. RSC Adv. 2018, 8, 38648–38655. [Google Scholar] [CrossRef] [PubMed]
  34. Liu, J.; Lee, J.Y.; Pan, L.; Obermyer, R.T.; Simizu, S.; Zande, B.; Li, J.; Sankar, S.G.; Johnson, J.K. Adsorption and Diffusion of Hydrogen in a New Metal-Organic Framework Material: [Zn(bdc)(ted)0.5]. J. Phys. Chem. C 2008, 112, 2911–2917. [Google Scholar] [CrossRef]
  35. Li, R.; Han, X.; Liu, Q.; Qian, A.; Zhu, F.; Hu, J.; Fan, J.; Shen, H.; Liu, J.; Pu, X.; et al. Enhancing Hydrogen Adsorption Capacity of Metal Organic Frameworks M(BDC)TED0.5 through Constructing a Bimetallic Structure. ACS Omega 2022, 7, 20081–20091. [Google Scholar] [CrossRef] [PubMed]
  36. Song, P.; Li, Y.; He, B.; Yang, J.; Zheng, J.; Li, X. Hydrogen Storage Properties of Two Pillared-Layer Ni(II) Metal-Organic Frameworks. Microporous Mesoporous Mater. 2011, 142, 208–213. [Google Scholar] [CrossRef]
  37. Kozlova, S.G.; Mirzaeva, I.V.; Ryzhikov, M.R. DABCO Molecule in the M2(C8H4O4)2·C6H12N2 (M = Co, Ni, Cu, Zn) Metal-Organic Frameworks. Coord. Chem. Rev. 2018, 376, 62–74. [Google Scholar] [CrossRef]
  38. Seki, K.; Mori, W. Syntheses and Characterization of Microporous Coordination Polymers with Open Frameworks. J. Phys. Chem. B 2002, 106, 1380–1385. [Google Scholar] [CrossRef]
  39. Wang, H.; Getzschmann, J.; Senkovska, I.; Kaskel, S. Structural Transformation and High Pressure Methane Adsorption of Co2(1,4-bdc)2dabco. Microporous Mesoporous Mater. 2008, 116, 653–657. [Google Scholar] [CrossRef]
  40. Senkovska, I.; Kaskel, S. High Pressure Methane Adsorption in the Metal-Organic Frameworks Cu3(btc)2, Zn2(bdc)2dabco, and Cr3F(H2O)2O(bdc)3. Microporous Mesoporous Mater. 2008, 112, 108–115. [Google Scholar] [CrossRef]
  41. Lee, J.Y.; Pan, L.; Huang, X.; Emge, T.J.; Li, J. A Systematic Approach to Building Highly Porous, Noninterpenetrating Metal-Organic Frameworks with a Large Capacity for Adsorbing H2 and CH4. Adv. Funct. Mater. 2011, 21, 993–998. [Google Scholar] [CrossRef]
  42. Senthil Raja, D.; Pan, C.C.; Chen, C.W.; Kang, Y.H.; Chen, J.J.; Lin, C.H. Synthesis of Mixed Ligand and Pillared Paddlewheel MOFs Using Waste Polyethylene Terephthalate Material as Sustainable Ligand Source. Microporous Mesoporous Mater. 2016, 231, 186–191. [Google Scholar] [CrossRef]
  43. Wu, H.; Zhou, W.; Yildirim, T. High-Capacity Methane Storage in Metal—Organic Frameworks M2(dhtp): The Important Role of Open Metal Sites. J. Am. Chem. Soc. 2009, 131, 4995–5000. [Google Scholar] [CrossRef] [PubMed]
  44. Li, D.Z.; Chen, L.; Liu, G.; Yuan, Z.Y.; Li, B.F.; Zhang, X.; Wei, J.Q. Porous Metal-Organic Frameworks for Methane Storage and Capture: Status and Challenges. New Carbon Mater. 2021, 36, 468–496. [Google Scholar] [CrossRef]
  45. Hu, T.D.; Jiang, Y.; Ding, Y.H. Computational Screening of Metal-Substituted HKUST-1 Catalysts for Chemical Fixation of Carbon Dioxide into Epoxides. J. Mater. Chem. A 2019, 7, 14825–14834. [Google Scholar] [CrossRef]
  46. Jiang, Y.; Hu, T.D.; Yu, L.Y.; Ding, Y.H. A More Effective Catalysis of the CO2 Fixation with Aziridines: Computational Screening of Metal-Substituted HKUST-1. Nanoscale Adv. 2021, 3, 4079–4088. [Google Scholar] [CrossRef] [PubMed]
  47. Zong, S.; Zhang, Y.; Lu, N.; Ma, P.; Wang, J.; Shi, X.R. A DFT Screening of M-HKUST-1 MOFs for Nitrogen-Containing Compounds Adsorption. Nanomaterials 2018, 8, 958. [Google Scholar] [CrossRef] [PubMed]
  48. Goings, J.J.; Ohlsen, S.M.; Blaisdell, K.M.; Schofield, D.P. Sorption of H2 to Open Metal Sites in a Metal-Organic Framework: A Symmetry-Adapted Perturbation Theory Analysis. J. Phys. Chem. A 2014, 118, 7411–7417. [Google Scholar] [CrossRef] [PubMed]
  49. Avci, G.; Velioglu, S.; Keskin, S. In Silico Design of Metal Organic Frameworks with Enhanced CO2 Separation Performances: Role of Metal Sites. J. Phys. Chem. C 2019, 123, 28255–28265. [Google Scholar] [CrossRef]
  50. Fu, Z.; Yi, J.; Chen, Y.; Liao, S.; Guo, N.; Dai, J.; Yang, G.; Lian, Y.; Wu, X. From Interwoven to Noninterpenetration: Crystal Structural Motifs of Two New Manganese-Organic Frameworks Mediated by the Substituted Group of the Bridging Ligand. Eur. J. Inorg. Chem. 2008, 2008, 628–634. [Google Scholar] [CrossRef]
  51. Lestari, W.W.; Ni’Maturrohmah, D.; Putra, R.; Suwarno, H.; Arrozi, U.S.F. Enhanced Hydrogen Sorption Properties over Mg2+ Modified Solvothermal Synthesized HKUST-1 (Mg2+/HKUST-1). IOP Conf. Ser. Mater. Sci. Eng. 2019, 509, 012152. [Google Scholar] [CrossRef]
  52. Świrk, K.; Delahay, G.; Zaki, A.; Adil, K.; Cadiau, A. Facile Modifications of HKUST-1 by V, Nb and Mn for Low-Temperature Selective Catalytic Reduction of Nitrogen Oxides by NH3. Catal. Today 2022, 384–386, 25–32. [Google Scholar] [CrossRef]
  53. Zhang, Z.; Zhang, L.; Wojtas, L.; Eddaoudi, M.; Zaworotko, M.J. Template-Directed Synthesis of Nets Based upon Octahemioctahedral Cages That Encapsulate Catalytically Active Metalloporphyrins. J. Am. Chem. Soc. 2012, 134, 928–933. [Google Scholar] [CrossRef]
  54. Huynh, N.T.X.; Na, O.M.; Chihaia, V.; Son, D.N. A Computational Approach towards Understanding Hydrogen Gas Adsorption in Co-MIL-88A. RSC Adv. 2017, 7, 39583–39593. [Google Scholar] [CrossRef]
  55. Vinh, N.Q.; Duyen, N.T.M.; Tran, N.L.B.; Van Nghia, N.; Vien, L.T.T.V.; Thanh, H.T.M.; Huynh, N.T.X. Computational Study on Enhancing SO2 Capture Capacity of M2(BDC)2TED (M = Mg, V, Co, or Ni). Quy Nhon Univ. J. Sci. 2024, 18, 91–100. [Google Scholar]
  56. Huynh, N.T.X.; Chihaia, V.; Son, D.N. Enhancing Hydrogen Storage by Metal Substitution in MIL-88A Metal-Organic Framework. Adsorption 2020, 26, 509–519. [Google Scholar] [CrossRef]
  57. Dubbeldam, D.; Calero, S.; Ellis, D.E.; Snurr, R.Q. RASPA: Molecular Simulation Software for Adsorption and Diffusion in Flexible Nanoporous Materials. Mol. Simul. 2016, 42, 81–101. [Google Scholar] [CrossRef]
  58. Lorentz, H.A. Ueber Die Anwendung Des Satzes Vom Virial in Der Kinetischen Theorie Der Gase. Ann. Phys. 1881, 248, 127–136. [Google Scholar] [CrossRef]
  59. Berthelot, D. Sur Le Mélange Des Gaz. Acad. Sci. 1898, 126, 1703–1855. [Google Scholar]
  60. Martin, M.G.; Siepmann, J.I. Transferable Potentials for Phase Equilibria. 1. United-Atom Description of n-Alkanes. J. Phys. Chem. B 1998, 102, 2569–2577. [Google Scholar] [CrossRef]
  61. Lyubchyk, A.; Esteves, I.A.A.C.; Cruz, F.J.A.L.; Mota, J.P.B. Experimental and Theoretical Studies of Supercritical Methane Adsorption in the Mil-53(Al) Metal Organic Framework. J. Phys. Chem. C 2011, 115, 20628–20638. [Google Scholar] [CrossRef]
  62. Dubbeldam, D.; Calero, S.; Vlugt, T.J.H.; Krishna, R.; Maesen, T.L.M.; Smit, B. United Atom Force Field for Alkanes in Nanoporous Materials. J. Phys. Chem. B 2004, 108, 12301–12313. [Google Scholar] [CrossRef]
  63. Dubbeldam, D.; Calero, S.; Vlugt, T.J.H.; Krishna, R.; Maesen, T.L.M.; Beerdsen, E.; Smit, B. Force Field Parametrization through Fitting on Inflection Points in Isotherms. Phys. Rev. Lett. 2004, 93, 088302. [Google Scholar] [CrossRef]
  64. Klimeš, J.; Bowler, D.R.; Michaelides, A. Chemical Accuracy for the Van der Waals Density Functional. J. Phys. Condens. Matter 2010, 22, 022201. [Google Scholar] [CrossRef] [PubMed]
  65. Klimeš, J.; Bowler, D.R.; Michaelides, A. Van der Waals Density Functionals Applied to Solids. Phys. Rev. B—Condens. Matter Mater. Phys. 2011, 83, 195131. [Google Scholar] [CrossRef]
  66. Dion, M.; Rydberg, H.; Schröder, E.; Langreth, D.C.; Lundqvist, B.I. Van der Waals Density Functional for General Geometries. Phys. Rev. Lett. 2004, 92, 246401. [Google Scholar] [CrossRef]
  67. Kresse, G.; Furthmüller, J. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  68. Kresse, G.; Furthmüller, J. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  69. Joubert, D. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B—Condens. Matter Mater. Phys. 1999, 59, 1758–1775. [Google Scholar]
  70. Blöchl, P.E. Projector Augmented-Wave Method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  71. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [PubMed]
  72. Pack, J.D.; Monkhorst, H.J. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  73. Henkelman, G.; Arnaldsson, A.; Jónsson, H. A Fast and Robust Algorithm for Bader Decomposition of Charge Density. Comput. Mater. Sci. 2006, 36, 354–360. [Google Scholar] [CrossRef]
  74. Sanville, E.; Kenny, S.D.; Smith, R.; Henkelman, G. Improved Grid-Based Algorithm for Bader Charge Allocation. J. Comput. Chem. 2007, 28, 899–908. [Google Scholar] [CrossRef] [PubMed]
  75. Xu, F.; Wu, Y.; Wu, J.; Lv, D.; Yan, J.; Wang, X.; Chen, X.; Liu, Z.; Peng, J. A Microporous Zn(bdc)(ted)0.5 with Super High Ethane Uptake for Efficient Selective Adsorption and Separation of Light Hydrocarbons. Molecules 2023, 28, 6000. [Google Scholar] [CrossRef] [PubMed]
  76. Guerrero-Medina, J.; Mass-González, G.; Pacheco-Londoño, L.; Hernández-Rivera, S.P.; Fu, R.; Hernández-Maldonado, A.J.; Hern, A.J.; Hern, S.P. Long and Local Range Structural Changes in M[(bdc)(ted)0.5] (M = Zn, Ni or Cu) Metal Organic Frameworks upon Spontaneous Thermal Dispersion of LiCl and Adsorption of Carbon Dioxide. Microporous Mesoporous Mater. 2015, 212, 8–17. [Google Scholar] [CrossRef]
  77. Mason, J.A.; Oktawiec, J.; Taylor, M.K.; Hudson, M.R.; Rodriguez, J.; Bachman, J.E.; Gonzalez, M.I.; Cervellino, A.; Guagliardi, A.; Brown, C.M.; et al. Methane Storage in Flexible Metal–Organic Frameworks with Intrinsic Thermal Management. Nature 2015, 527, 357–361. [Google Scholar] [CrossRef] [PubMed]
  78. Lee, K.; Howe, J.D.; Lin, L.C.; Smit, B.; Neaton, J.B. Small-Molecule Adsorption in Open-Site Metal-Organic Frameworks: A Systematic Density Functional Theory Study for Rational Design. Chem. Mater. 2015, 27, 668–678. [Google Scholar] [CrossRef]
  79. Vlaisavljevich, B.; Huck, J.; Hulvey, Z.; Lee, K.; Mason, J.A.; Neaton, J.B.; Long, J.R.; Brown, C.M.; Alfè, D.; Michaelides, A.; et al. Performance of van Der Waals Corrected Functionals for Guest Adsorption in the M2(dobdc) Metal-Organic Frameworks. J. Phys. Chem. A 2017, 121, 4139–4151. [Google Scholar] [CrossRef]
Figure 1. The configuration of (a) a primitive unit cell and (b) a three-dimensional view of M2(BDC)2(DABCO) or M(DABCO).
Figure 1. The configuration of (a) a primitive unit cell and (b) a three-dimensional view of M2(BDC)2(DABCO) or M(DABCO).
Crystals 14 00596 g001
Figure 2. (a) The simulated N2 adsorption isotherms of the M(DABCO) at 77 K and (b) the corresponding pore size distribution (PSD) of the M(DABCO).
Figure 2. (a) The simulated N2 adsorption isotherms of the M(DABCO) at 77 K and (b) the corresponding pore size distribution (PSD) of the M(DABCO).
Crystals 14 00596 g002
Figure 3. The simulated CH4 adsorption isotherms of M(DABCO) with the total uptake (solid lines) and the excess uptake (dashed lines) at 298 K and pressures under 1 bar: (a) gravimetric adsorption and (b) volumetric adsorption.
Figure 3. The simulated CH4 adsorption isotherms of M(DABCO) with the total uptake (solid lines) and the excess uptake (dashed lines) at 298 K and pressures under 1 bar: (a) gravimetric adsorption and (b) volumetric adsorption.
Crystals 14 00596 g003
Figure 4. The simulated CH4 adsorption isotherms of M(DABCO) with the total uptake (solid lines) and the excess uptake (dashed lines) at 298 K and pressures up to 100 bar: (a) gravimetric adsorption and (b) volumetric adsorption.
Figure 4. The simulated CH4 adsorption isotherms of M(DABCO) with the total uptake (solid lines) and the excess uptake (dashed lines) at 298 K and pressures up to 100 bar: (a) gravimetric adsorption and (b) volumetric adsorption.
Crystals 14 00596 g004
Figure 5. (a) The simulated total volumetric CH4 sorption isotherms of M(DABCO) at 238 K–298 K and pressures up to 100 bar and (b) the dependence of the simulated volumetric CH4 uptakes of M(DABCO) on the temperature.
Figure 5. (a) The simulated total volumetric CH4 sorption isotherms of M(DABCO) at 238 K–298 K and pressures up to 100 bar and (b) the dependence of the simulated volumetric CH4 uptakes of M(DABCO) on the temperature.
Crystals 14 00596 g005
Figure 6. The dependence of the simulated maximum total and excess CH4 uptakes (at 298 K) on (a) surface area and (b) pore volume (solid and empty circles). Herein, solid lines denote the linear fitting with the maximum uptakes.
Figure 6. The dependence of the simulated maximum total and excess CH4 uptakes (at 298 K) on (a) surface area and (b) pore volume (solid and empty circles). Herein, solid lines denote the linear fitting with the maximum uptakes.
Crystals 14 00596 g006
Figure 7. The simulated adsorption heat (Qst) of CH4 on M(DABCO). Inset: calculated Qst at low pressures below 1 bar.
Figure 7. The simulated adsorption heat (Qst) of CH4 on M(DABCO). Inset: calculated Qst at low pressures below 1 bar.
Crystals 14 00596 g007
Figure 8. DOS of the adsorbed CH4 molecule on M(DABCO) compared to that of the free CH4 molecule: (a) on adsorption site 1 (the M-O-C cluster) and (b) on adsorption site 2 (the interace between the M-O-C cluster and TED).
Figure 8. DOS of the adsorbed CH4 molecule on M(DABCO) compared to that of the free CH4 molecule: (a) on adsorption site 1 (the M-O-C cluster) and (b) on adsorption site 2 (the interace between the M-O-C cluster and TED).
Crystals 14 00596 g008
Table 1. Lennard-Jones parameters (ε, σ) of the atoms of M(DABCO).
Table 1. Lennard-Jones parameters (ε, σ) of the atoms of M(DABCO).
Atoms ε / k B (K) σ (Å)
H a7.652.85
C a47.863.47
N a38.953.26
O a48.163.03
Mg b55.862.69
Mn b6.542.64
Fe b6.542.59
Co b7.052.56
Ni b7.552.52
Cu b2.523.11
Zn b62.402.46
CH4 (Dubbeldam et al.) [62,63]158.50 3.72
CH4 (TraPPE-UA) (Martin et al.) [60] 148.00 3.73
a DREIDING force field; b the universal force field (UFF).
Table 2. The simulated geometrical features, including specific surface area, pore volumes, and pore sizes of M(DABCO), compared to the available simulated and experimental data.
Table 2. The simulated geometrical features, including specific surface area, pore volumes, and pore sizes of M(DABCO), compared to the available simulated and experimental data.
MOFs A B E T (m2/g)Vp (cm3/g)Pore Size (Å)
Cu(DABCO)15610.718.53
Fe(DABCO)15710.728.45
Zn(DABCO)15970.738.72
Co(DABCO)16280.748.61
Ni(DABCO)16860.768.69
Mn(DABCO)17010.778.71
Mg(DABCO)19310.878.72
Cu(DABCO) 1631 [30], 1572 [76]0.63 [30], 0.65 [76]7.98 [76]
Zn(DABCO)1781 [30], 1523 [76],
1904 [75]
0.65 [30], 0.56 [76],
0.73 [75]

8.00 [75]
Co(DABCO)1708 [30], 1600 [39]0.62 [30], 0.82 [39]
Ni(DABCO) 1905 [30], 1698 [76]0.76 [30], 0.71 [76]7.88 [76]
HKUST-1 [6]18500.78
Table 3. The simulated maximum gravimetric and volumetric CH4 adsorption on M(DABCO) for the total uptake (at 100 bar) and excess uptake (at 60 bar for M = Cu, Fe, Zn and 65 bar for M = Co, Ni, Mn, Mg), compared with available highest recorded data for CH4 adsorption of MOFs.
Table 3. The simulated maximum gravimetric and volumetric CH4 adsorption on M(DABCO) for the total uptake (at 100 bar) and excess uptake (at 60 bar for M = Cu, Fe, Zn and 65 bar for M = Co, Ni, Mn, Mg), compared with available highest recorded data for CH4 adsorption of MOFs.
MOFsCH4 Adsorption Heat   Adsorption ,   Q s t = H a d s (kJ/mol)
Gravimetric UptakeVolumetric Uptake
cc(STP)/cc
mg/gcm3/g
m t o t n e x c m t o t m e x c V t o t V e x c
Cu(DABCO)182.39135.12254.83188.79220.26163.1717.14
Fe(DABCO)186.58138.71260.69193.80222.14165.1417.19
Zn(DABCO)189.59141.12264.89197.17226.30168.4417.21
Co(DABCO)194.05143.93271.12201.09227.03168.3917.17
Ni(DABCO)199.90148.48279.30207.46228.68169.8617.10
Mn(DABCO)202.22150.06282.53209.65228.94169.8917.13
Mg(DABCO)231.39172.30323.29240.73231.43172.3317.27
Co(DABCO) [39]14 wt.% (75 bar) a
Cu(DABCO) [38] ca. 185 (35 bar)188 (35 bar) a16.25
HKUST-1 [6] 216 (184) b178 (165) c 267 (227) b220 (204) b17
PCN-14 [6]197 (169) b157 (146) c 230 (195) b183 (171) b18.7
MIL-101(Cr) 215 (150) b
Al-soc-MOF-1 [13,19]420 (263) b -579 (362) b 197 (123) b 10.5
NJU-Bai 43 [26] 396 (315) b 254 (202) b 14.45
UTSA-76 [27]263 (216) b 302 (257) b 257 (211) b 15.44
UTSA-110a [8] 402 (312) b 241 (187) b 15.49
DOE500 263
a at a temperature of 303 K; b at a temperature of 298 K and a pressure of 65 (35) bar.
Table 4. The simulated total and excess volumetric uptakes (cc/cc at STP) of CH4 on M(DABCO) at other pressures and temperatures.
Table 4. The simulated total and excess volumetric uptakes (cc/cc at STP) of CH4 on M(DABCO) at other pressures and temperatures.
Temperature (K) V t o t V e x c
(Max)
35 bar65 bar100 bar (Max)
238270.61284.95292.15242.58 (30.0 bar)
248257.89274.09282.54230.32 (35.0 bar)
258243.37263.09274.05218.13 (37.5 bar)
268228.56250.63263.34206.24 (42.5 bar)
278214.01238.87252.54194.15 (47.5 bar)
288199.34226.72242.40183.14 (52.5 bar)
298184.99214.38231.43172.71 (60.0 bar)
Table 5. The vdW-DF-based calculated adsorption energy ( E a d s ) of CH4 on sites close to the metals of M(DABCO) and the nearest distance between the methane molecule and M(DABCO).
Table 5. The vdW-DF-based calculated adsorption energy ( E a d s ) of CH4 on sites close to the metals of M(DABCO) and the nearest distance between the methane molecule and M(DABCO).
Adsorption ConfigurationsSite 1
(Metal Cluster)
Site 2
(M-O-C Cluster—TED Interface)
E a d s
(kJ/mol)
d 1
(Å)
E a d s
(kJ/mol)
d 2
(Å)
Mg−76.582.92−76.232.46
Mn−77.733.08−76.152.58
Fe−76.973.00−75.782.50
Co−76.562.97−76.552.59
Ni−73.353.01−75.402.46
Cu−20.312.87−45.332.74
Zn−61.462.89−62.132.49
Table 6. The Bader charges of the CH4 molecule sorbed on M(DABCO) MOFs (in e) compared those of the isolated corresponding compounds on the adsorption site of the metal cluster (site 1).
Table 6. The Bader charges of the CH4 molecule sorbed on M(DABCO) MOFs (in e) compared those of the isolated corresponding compounds on the adsorption site of the metal cluster (site 1).
CH4@M(DABCO) at Site 1M = MgM = MnM = FeM = CoM = NiM = CuM = Zn
CH4 molecule4 H+0.011−0.011+0.037−0.028−0.011−0.007−0.001
1 C−0.018+0.005−0.030+0.016+0.001+0.012−0.006
Total−0.007−0.006+0.007−0.012−0.010−0.011−0.007
M(DABCO)20 H−0.008+0.053−0.043+0.020−0.053−0.007+0.013
12 C1 −0.052+0.021+0.017+0.035+0.151+0.165−0.091
4 C2 +0.081+0.024+0.023−0.029−0.160−0.178+0.108
6 C3−0.016−0.056+0.016−0.021−0.003−0.034+0.033
2 N+0.014+0.011−0.0004−0.010+0.006−0.002−0.019
8 O−0.013−0.036−0.010+0.059+0.066+0.099−0.043
2 M+0.001−0.012−0.009−0.042+0.003−0.035+0.004
Total+0.007+0.006−0.006+0.012+0.010+0.011+0.007
Table 7. The Bader charges of the CH4 molecule sorbed on M(DABCO) MOFs (in e) compared those of the isolated corresponding compounds on the adsorption site of the bridge between the metal cluster and the TED (site 2).
Table 7. The Bader charges of the CH4 molecule sorbed on M(DABCO) MOFs (in e) compared those of the isolated corresponding compounds on the adsorption site of the bridge between the metal cluster and the TED (site 2).
CH4@M(DABCO) at Site 2M = MgM = MnM = FeM = CoM = NiM = CuM = Zn
CH4 molecule4 H+0.032+0.022+0.033−0.004+0.028+0.022−0.043
1 C−0.038−0.027−0.024−0.001−0.036−0.031+0.036
Total−0.006−0.005+0.009−0.005−0.008−0.009−0.007
M(DABCO)20 H+0.060+0.008−0.077+0.061+0.009+0.034+0.020
12 C1 −0.044+0.037+0.069−0.016+0.130+0.096−0.089
4 C2 −0.011+0.007−0.015−0.023−0.170−0.162+0.111
6 C3−0.032−0.020+0.0003−0.042−0.022−0.001+0.023
2 N+0.016+0.015+0.006+0.002+0.006+0.032−0.034
8 O+0.016−0.034+0.012+0.047+0.054+0.042−0.054
2 M+0.001−0.007−0.001−0.025+0.001−0.032+0.031
Total+0.006+0.005−0.006+0.005+0.008+0.009+0.007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Huynh, N.T.X.; Nguyen-Van, T.; Tran, N.L.B.; Nghia, N.V.; Thanh, P.N. Exploring Methane Storage Capacities of M2(BDC)2(DABCO) Sorbents: A Multiscale Computational Study. Crystals 2024, 14, 596. https://doi.org/10.3390/cryst14070596

AMA Style

Huynh NTX, Nguyen-Van T, Tran NLB, Nghia NV, Thanh PN. Exploring Methane Storage Capacities of M2(BDC)2(DABCO) Sorbents: A Multiscale Computational Study. Crystals. 2024; 14(7):596. https://doi.org/10.3390/cryst14070596

Chicago/Turabian Style

Huynh, Nguyen Thi Xuan, Tue Nguyen-Van, Nguyen Le Bao Tran, Nguyen Van Nghia, and Pham Ngoc Thanh. 2024. "Exploring Methane Storage Capacities of M2(BDC)2(DABCO) Sorbents: A Multiscale Computational Study" Crystals 14, no. 7: 596. https://doi.org/10.3390/cryst14070596

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop