Next Article in Journal
circSMARCA5 Is an Upstream Regulator of the Expression of miR-126-3p, miR-515-5p, and Their mRNA Targets, Insulin-like Growth Factor Binding Protein 2 (IGFBP2) and NRAS Proto-Oncogene, GTPase (NRAS) in Glioblastoma
Next Article in Special Issue
Influence of the Substituent’s Size in the Phosphinate Group on the Conformational Possibilities of Ferrocenylbisphosphinic Acids in the Design of Coordination Polymers and Metal–Organic Frameworks
Previous Article in Journal
Sex- and Neuropsychiatric-Dependent Circadian Alterations in Daily Voluntary Physical Activity Engagement and Patterns in Aged 3xTg-AD Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Silico Screening of Metal-Organic Frameworks for Formaldehyde Capture with and without Humidity by Molecular Simulation

1
Energy and Electricity Research Center, Jinan University, Zhuhai 519070, China
2
School of Energy and Power Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(22), 13672; https://doi.org/10.3390/ijms232213672
Submission received: 10 October 2022 / Revised: 4 November 2022 / Accepted: 6 November 2022 / Published: 8 November 2022
(This article belongs to the Special Issue Properties and Applications of Metal-Organic Frameworks)

Abstract

:
Capturing formaldehydes (HCHO) from indoor air with porous adsorbents still faces challenges due to their low capacity and poor selectivity. Metal-organic frameworks (MOFs) with tunable pore properties were regarded as promising adsorbents for HCHO removal. However, the water presence in humid air heavily influences the formaldehyde capture performance due to the competition adsorption. To find suitable MOFs for formaldehyde capture and explore the relationship between MOFs structure and performance both in dry air and humid air, we performed grand canonical Monte Carlo (GCMC) molecular simulations to obtain working capacity and selectivity that evaluated the HCHO capture performance of MOFs without humidity. The results reveal that small pore size (~5 Å) and moderate heat of adsorption (40–50 kJ/mol) are favored for HCHO capture without water. It was found that the structure with a 3D cage instead of a 2D channel benefits the HCHO adsorption. Atoms in these high-performing MOFs should possess relatively small charges, and large Lennard-jones parameters were also preferred. Furthermore, it was indicated that Henry’s constant (KH) can reflect the HCHO adsorption performance without humidity, in which the optimal range is 10−2–101. Hence, Henry’s constant selectivity of HCHO over water (SKH HCHO/H2O) and HCHO over mixture components (H2O, N2, and O2) was obtained to screen MOFs at an 80% humidity condition. It was suggested that SKH for the mixture component overestimates the influence of N2 and O2, in which the top structures absorb a quantity of water in GCMC simulation, while SKH HCHO/H2O can efficiently find high-performing MOFs for HCHO capture at humidity in low adsorption pressure. The ECATAT found in this work has 0.64 mol/kg working capacity, and barely adsorbs water during 0–1 bar, which is the promising candidate MOF for HCHO capture.

1. Introduction

Volatile organic compounds (VOCs) include a variety of chemicals, some of which may have short- and long-term adverse health effects. Among the most popular VOCs, formaldehyde (HCHO), is very allergenic and carcinogenic even at very low concentrations [1,2,3]. The World Health Organization (WHO) recommended that a safe concentration of formaldehyde vapor for humans must be below 0.08 ppm (30-min) and a threshold sensory irritation of 0.1 mg/m3, which also can be lethal at a concentration of 30 mg/m [4,5,6]. Therefore, the removal of formaldehyde from contaminated air or the industrial process is in demand. The various methods for abating HCHO from indoor areas include photocatalysis [7,8], catalytic oxidation [9,10], and adsorption-based [11,12], and formaldehyde capture has been one of the most promising strategies due to the facile operation [13]. To date, a number of adsorbents including activated carbons [14], zeolites [13], SiO2 [15,16], AlOOH [17], amine-supported materials [18], etc. have been explored for formaldehyde removal. However, these traditional adsorbents are not suitable for addressing the continuous release of formaldehyde due to their non-polarity and highly amorphous nature [6], especially for the low adsorption capacity and poor selectivity in humid air [14,19,20].
Metal-organic frameworks (MOFs), as an emerging class of porous crystalline materials composed of inorganic metal nodes and organic ligands, have attracted increasing research interest due to their high specific surface areas and pore volumes, well-defined porosity, and tunable pore properties, which have been utilized for gas storage, separation, and catalysis [21,22]. Many MOFs (ZIF-67 [23], ZIF-8 [24], UiO-66 [25]) exhibited strong affinity toward HCHO and were reported as chemical capacitance sensors to detect it with satisfactory responses at room temperature. The potential for removal of HCHO was also demonstrated from a number of case studies on diverse MOFs. Wang et al. [26] prepared α, β, γ-CD-MOFs to capture HCHO at room temperature, and found that γ-CD-MOFs can almost totally remove 0.5 mg/m3 HCHO within 15 min, which indicated the high efficiency of the MOFs adsorbent even in such a low concentration. Accordingly, diamine-appended MIL-101(Cr) with water resistibility was synthesized, and it was found that the highest adsorption capacity can reach 5.4 mol/kg in 150 ppm concentration. [27] Nevertheless, the strong competitive of water molecules in humid indoor air hinders the adsorption of HCHO by occupying the adsorption sites preferentially. It was noted that the formaldehyde capacity of UiO-66-NH2 from 27.67 mg/g without humidity decreased to 9.84 mg/g at 12% relative humidity [28]. It is essential to obtain the adsorption performance with and without humidity to investigate the competitive adsorption between water and the HCHO of MOFs.
The arrangement of metal nodes and organic linkers, as well as the self-assembly topology of structures, leads to a diversity of experimental MOFs [29]. Moreover, there are also millions of hypothetical MOFs (hMOFs) by computational design [30], which makes it extremely difficult to search for target MOFs applied in HCHO removal. High-throughput computational screening (HTCS) based on the grand canonical Monte Carlo (GCMC) has become an efficient strategy to quickly find suitable MOFs for adsorption and separation applications, including carbon capture [31], H2 storage [32], and H2S separation [22]. During the screening, the structure-property relationship can be extracted to guide the rational design of high-performance porous adsorbents. Our previous work [33] screened out Y-BTC, ZnCar, and Ni-BIC from 2932 kinds of Computational-ready, experimental (CoRE) MOFs, and indicated that it has better capture and regeneration performance compared with activated carbon in a high HCHO concentration. However, the correlation between the adsorption performance and MOF characteristics remains to be clarified, especially the inside information of atoms in high-performing MOFs. Regarding the competition adsorption between water and HCHO, Yuan et al. [12] recently identified hydrophilic and hydrophobic MOFs by the Henry’s constant (KH), referring to the value of ZIF-8. They regarded MOFs with KH < 2.6 × 106 mmol/(g·Pa) as hydrophobic and then evaluated the adsorption capacity and selectivity of 31,399 hMOFs without humidity by GCMC and machine learning. Hence, the competition adsorption behavior between water and HCHO, as well as the structure-property relationship for MOFs with humidity, await further investigation.
Considering the HCHO capture performance with and without humidity focus in this work, the capacity and selectivity of HCHO in N2 and O2 (dry air) were established and the structure-property relationship was extracted in dry air, including the charge and Lennard-jones parameter of atoms in MOFs. The KH selectivity (SKH) was then used to screen high-performance MOFs in humid air, in which the adsorption isotherm at the ambient temperature was simulated to verify it. It was suggested that SKH of HCHO over water can play a crucial role in screening suitable MOFs for HCHO capture in low pressure. This work offers the molecular understanding of the rational design of high-performing MOFs for HCHO removal both in dry and humid air.

2. Results and Discussion

2.1. HCHO Capture Performance without Humidity

In this work, GCMC simulation was carried out to evaluate the HCHO pressure swing adsorption (PSA) performance of 1668 CoRE-MOFs in dry air between 0.1 and 1 bar. Figure 1 shows the relationship of working capacity (ΔW), selectivity (S), and heat of adsorption (Qst) to the pore size. It is revealed that the highest working capacity is 4.01 mol/kg, the refcode in the Cambridge Structural Database (CSD) is LAVSUY, followed by the DUBWON (3.98 mol/kg), and PARMIG (3.93 mol/kg). It denotes LCD located in the range of 5–6 Å for most MOFs with ΔW > 2 mol/kg and selectivity over 103, presented in Figure 1a. Notably, Bellat et al. [13]. previously reported the total uptake of 2.3 mol/kg (7 wt%) of Ga-MIL-53 at 2000 ppm, room temperature, and Wang et al. [27] found the adsorption uptake of 3.34 mol/kg in MIL-101(Cr) at 150 ppm and then can up to 5.49 mol/kg after being post-modified by ethylenediamine. The highest working capacity in this work is larger than the capacity of MOFs without modification in the experiment, which indicated that there are promising MOFs awaiting discovery. In addition, most MOFs with large working capacities (>2 mol/kg) have relatively small LCDs (4–6 Å), which is nearly double the formaldehyde dynamic diameter (2.43 Å) [34]. This may be ascribed to the suitable interaction of MOFs in such a pore size, which reflected in the moderate heat of desorption (40~50 kJ/mol). A similar phenomenon of pore size dependence for gas adsorption also can be found in Wilmer et al.’s [35]. CH4 storage work and Banerjee et al.’s [36] Xe/Kr separation work. Moreover, Figure 1b indicated that the selectivity fluctuates with the heat of desorption, while higher Qst benefits from the increase of selectivity. When the LCD > 10 Å, it was found that most structures exhibit poor HCHO capture performance (ΔW < 2.0 mol/kg, S < 103 and Qst < 40 kJ/mol). However, as presented in Figure S1, it was noted that excessive strong interaction leads to the HCHO being difficult to desorb during the PSA process, which will be discussed later.
In indoor air, formaldehyde is generally found in trace amounts with very low partial pressure, which adsorbs in Henry’s law region. Thus, Figure 2a presented the correlation between adsorption performance and KH. It was found that the MOFs exhibited poor ΔW and S when KH was small than 10−2 mol/(kg·Pa), which can be ascribed to the poor interaction that limited the HCHO adsorption. As for those structures with KH > 101 mol/(kg·Pa), it is suggested that strong interaction benefits the selectivity of HCHO over N2 and O2. However, as shown in Figure S2, the overlarge host-adsorbate interaction causes the HCHO extremely hard to desorb, which is not conducive to HCHO capture. It was indicated that 10−2 ≤ KH ≤ 101 was favored for high working capacity and moderate selectivity that benefits HCHO capture. Such a result suggests that KH can be used to pre-screen adsorbents for HCHO capture in dry air. Moreover, the correlation between atom distribution in crystal (MaxER, MinER) and ΔW was also investigated. Notably, as shown in Figure S3a, the MaxER = MinER = 0.33 indicated that the MOF crystal is cubic topology and x, y, z axisymmetric. For MinER = 0, the structures tend to be 2-D layers. Hence, when the MinER is close to 0, it means the pore of structures tends to be a channel instead of a cage. In Figure 2b, it was found that MaxER near 0.4 and MinER located in 0.2–0.3 are favored for high working capacity, while most structures with MaxER > 0.6 and MinER < 0.1 exhibit low working capacity. Compared with the channel, 3D cages with different properties in each direction were preferred in HCHO capture.
We further analyzed the correlation between adsorption performance and the force field of atoms in MOFs, including average positive/negative charge(APC/ANC) and LJ parameters. According to Figure 3a, APC < 0.2 (ANC > −0.2) and APC > 0.5 (ANC < −0.5) exhibit poor working capacity, which the APC located in 0.2–0.5 for most structures with ΔW > 0.2 mol/kg. As shown in Figure S4a, it was suggested that the enhancement of LJ interaction always favors the increase in HCHO capacity. The excessive interaction makes the HCHO difficult to desorb from the MOFs in Figure S4b, which led to a decrease in working capacity, similar to the tendency found in KH. Furthermore, as shown in Figure 3b, the selectivity of HCHO over N2 and O2 scatter in a wide range (1–107), can be divided into three parts according to the LJ parameters. For those MOFs with 0 < Aε < 1.5 and 0 < Aσ < 0.12, they have large enough pore volume (Figure S4c), but the weak interaction limited the adsorption of HCHO, which makes the selectivity lower than 104. As for those structures with Aε > 3 and Aσ > 0.21, the tiny pore volume cannot afford higher capacity, which also makes it unsuitable for HCHO capture. Therefore, it was suggested that 1.5 ≤ Aε ≤ 3.0 and 0.12 ≤ Aσ ≤ 0.21 are beneficial to the enhancement of selectivity, and most structures with S > 105 are located in this range. Moreover, it was also found that the high selectivity accompanied by satisfying capacity in Figure S4d, which indicated the combined moderate charge and LJ parameters, favors the HCHO capture performance of MOFs.
The top 10 MOFs with excellent formaldehyde capture performance are listed in Table 1. Among them, the best MOF is LAVSUY, with 6.62 Å LCD, 0.43 MaxER, 0.50 e APC, 0.16 kcal/mol Aε, 1.18 × 10−2 mol/(kg·Pa) KH, which was predicted to have 4.01 mol/kg working capacity and 2722 selectivity. As shown in Figure S3c, the LAVSUY has bcu (body-centered cubic) topology, Y nodes connected by 1,3,5-Benzenetricarboxylic acid. In addition, other top-performance MOFs exhibited similar structural characteristics. For example, LCD located in 4.25–6.62 Å, MaxER in 0.37–0.56, APC in 0.16–0.59, Aε in 0.13–0.23 kcal/mol, KH in 7.37 × 10−2–2.68 × 10−1 mol/(kg·Pa), and other descriptors are provided in Table S2, which is quite consistent with the suitable range for HCHO capture found in previous results.
The adsorption isotherm of HCHO, N2, and O2 mixture components obtained from GCMC simulation for the top 3 MOFs (LAVSUY, DUBWON, and PARMIG) are presented in Figure 4a–c. All MOFs almost were Type I adsorption isotherm [37] defined by IUPAC and exhibited ultra-high capacity with extremely low N2 and O2 capacity. It is worthy of note that the DUBWON and PARMIG seem to reach the saturation capacity when the pressure is larger than 0.8 bar, whereas the LAVSUY probably tends to have a higher capacity as the pressure continues to increase. Moreover, combined with the snapshots of Figure S3c–e, the density plots in Figure 4d–f illustrated that the HCHO majority adsorb in the center of the cage close to the metal nodes of MOFs, which is consistent with the results of Figure 2b.

2.2. HCHO Capture Performance with Humidity

As we mentioned before, the competitive adsorption between HCHO and H2O would heavily influence the HCHO capture performance in humid air. However, estimating the HCHO capture performance for a large quantity of MOFs via GCMC simulation or experiment is extremely time-consuming [38]. It was proposed that the KH of water can be adapted to identify whether the MOFs are hydrophilic or hydrophobic in HCHO capture. Moreover, the results in dry air suggested that KH are the dominant factor to determine the HCHO capture performance. Thus, regarding the heavy competition between water and HCHO, there are two Henry’s selectivity (SKH) were calculated to screen out suitable MOFs in humid air, type 1: HCHO over water, type 2: HCHO over water, N2, and O2. As shown in Figure 5a, it was found that the LCD of the top 3 MOFs for SKH HCHO/H2O is located in a wide range (5–13 Å). Whereas the small LCD (~5 Å) exhibited better performance for SKH HCHO/(H2O + N2 + O2) in Figure 5b, similar to the trend found in dry air.
Figure 6 was presented to illustrate the relationship between Henry constant selectivity and chemical descriptor, including MPC, MNC, Aσ, and Aε. It was found that the SKH HCHO/H2O depend significantly on the charge since they are nonpolar adsorbates. In Figure 6a, it was found that most SKH HCHO/H2O > 10 MOFs with MPC < 2. As for MPC ≥ 2, a large quantity of MOFs exhibited SKH HCHO/H2O < 10−2 due to the strong Coulombic interaction between MOFs and water. Moreover, as shown in Figure S5a,c, it was suggested that top MOFs for SKH HCHO/H2O exhibited low void fraction and high LJ descriptors (Aσ > 0.2 and Aε > 3), including ECAHAT (LCD~12 Å). Moreover, as shown in Figure 6b, high Aσ and Aε also benefit the increment of SKH HCHO/(H2O + N2 + O2), which indicated that Lennard-jones interaction is a dominant role in determining the HCHO capture performance in humid air. Moreover, in Figure S5b,d, it was found most MPC > 2 MOFs have extremely large KH for water that is not favored both in SKH HCHO/H2O and SKH HCHO/(H2O + N2 + O2).
The LCD, MPC, MNC, and KH of the top 10 MOFs for SKH HCHO/H2O were provided in Table 2. JAVTAC has a maximum SKH HCHO/H2O, which is 418.76, followed by WOJJOV (194.11), and ECAHAT (144.74). Notably, it was found that all the MOFs in Table 2 have extremely low KH of N2 and O2, which indicated they probably have a poor affinity toward N2 and O2. Indeed, as shown in Figure S6, SKH HCHO/(H2O + N2 + O2) is almost linear with the SKH HCHO/H2O for those MOFs with KH of water > 1. The ranking difference between SKH HCHO/H2O and SKH HCHO/(H2O + N2 + O2) majority are those MOFs with KH of water < 1.
In order to verify Henry’s constant screening results, the GCMC simulations were implemented for six MOFs (JAVTAC, WOJJOV, ECAHAT, DORDUK, DOTTUC, and OHOMIH) to obtain the adsorption isotherm under 80% humidity conditions in 298 K. As shown in Figure 7d–f, all of the structures selected by SKH HCHO/(H2O + N2 + O2) are highly hydrophilic structures that adsorb a lot of water (>4 mol/kg) in low pressure (0.1 bar). The JAVTAC has a higher formaldehyde uptake in lower pressure and is in agreement with Henry’s law. However, when the pressure gradually increases to 1.0 bar, the water molecules with polar functional groups occupy the adsorption sites preferentially, and the strong competitive adsorption of water molecules hinders the capture of HCHO [12]. The water uptake then exhibits an s-shaped isotherm and finally reaches a higher water loading in structures. As for WOJJOV, the water exhibited a similar trend with JAVTAC, while the HCHO uptake maintains at 0.9 mol/kg. For the ECAHAT, the water uptake is extremely low during the whole pressure range, and it has a 0.64 mol/kg working capacity between 0.1 bar and 1 bar, and 465 selectivity of HCHO over H2O, N2, and O2, which is a promising candidate for HCHO capture under humidity conditions. In this study, it was indicated that SKH HCHO/H2O can be recognized as a critical descriptor in low pressure. It remains a challenge to find a suitable descriptor for screening MOFs under humidity conditions in high pressure with reasonable computation cost.

3. Materials and Methods

3.1. MOFs Database

All MOF structures were obtained from the computation-ready, experimental (CoRE) MOF database Version 1.0 [39], which the solvent and disorder structures were removed from Cambridge structural database (CSD) by Chung and co-workers. The structure with density derived electrostatic and chemical (DDEC) [40] charges containing 2932 structures were developed by Nazarian [41]. After removing the structures with zero accessible surface area (ASA), there are 1668 structures to perform formaldehyde capture screening. The ASA, largest cavity diameter (LCD), pore limiting diameter (PLD), and available pore volume (Va) were computed using the 1.86 Å nitrogen probe in zeo++0.3. Helium void fraction (VF) and Henry’s constant of MOFs toward H2O, HCHO, N2 and O2 were obtained by the Widom particle insertion method.

3.2. Grand Canonical Monte Carlo

CoRE MOFs containing 1668 structures carrying DDEC charges were employed for high-throughput screening. GCMC simulations were implemented to obtain the adsorption performance of these structures in RASPA 2.0. During the screening stage, 4 × 104 Monte Carlo cycles were performed to estimate the adsorption isotherms of each MOF, including the initial 2 × 104 cycles of equilibration run, and the other 2 × 104 cycles of the production run. Four Monte Carlo moves of insertion, deletion, rotation, and translation were implemented with equal probability. Identity change of adsorbate molecules for multi-component adsorption was performed with the two-fold probability of insertion, deletion, rotation, and translation moves. The simulation temperature was maintained at 298 K, and the pressure ranges from 0.1 bar to 1 bar with a molar ratio of HCHO:N2:O2 = 2:798:200. As for the GCMC simulation under the 80% humidity condition, a total of 4 × 106 Monte Carlo cycles were carried out. The simulation of the molar ratio for HCHO:H2O:N2:O2 was set as 200/3280/77,216/19,304 to represent the 200 ppm HCHO concentration in the humid air. The adsorption performance, including working capacity (ΔW), selectivity (S) of HCHO in dry air, and KH selectivity (SKH) of HCHO in humid air were computed using the following equation. The heat of adsorption of HCHO was obtained at 200 Pa in the pure component.
Δ W = W HCHO , 1 bar W HCHO , 0.1 bar
S = W HCHO , 1 bar / f HCHO W i , 1 bar / f i
SK H = K H , HCHO / f HCHO K H , i / f i
where W HCHO , 1 bar is the HCHO capacity at 298 K, 1 bar, f i is the fraction of gas component (HCHO, H2O, N2 and O2) in the mixture adsorbate.

3.3. Force Field

During molecular simulation, the Lennard-Jones (LJ) and Coulomb potentials were used to describe the non-bonded interactions between MOFs and adsorbates.
V i j = 4 ε i j σ i j r i j 12 σ i j r i j 6 + q i q j 4 π ε 0 r i j
Herein, ij represents the two interacting atoms, where ε is the depth of the potential wall, σ i j is the finite distance at which the inter-particle potential is zero, r i j is the distance between the particles. All the LJ parameters of MOFs were taken from UFF force field [42], and the LJ parameters for N2 and O2 were adapted from the TraPPE force field [43]. The Lorentz-Berthelot mixing rule was applied for inter-atomic LJ interactions. q i and q j are the atomic partial charges of two interacting atoms, and ε 0 is the vacuum permittivity constant. Long-range Coulombic interaction was described by the Ewald method [44] with a cutoff of 12.8 Å. We calculate the average sigma of LJ interaction and the average epsilon of LJ interaction, represented as Aσ and Aε.
The water model we adapted is the Tip4p force field, for it can well represent the water adsorption property in hydrophobic MOFs [45]. The force field parameters of formaldehyde were taken from Hantal et al.’s study [46] in which the planar formaldehyde model was employed. The bond lengths of H-C and C=O are 1.101 and 1.203 Å, respectively, and the angle of H-C=O is 121.8°. In this model, only the C and O atoms carry fractional charges of +0.45 × 101 and −0.45 × 101, respectively, and a dipole moment of 2.6 D along C=O bond vector was applied. The N2 and O2 force field are taken from TraPPE. All of the parameters of adsorbates are summarized in Table S1.

3.4. The Descriptor of MOF Characteristic

There are nine descriptors that were collected from the crystallographic information file (CIF) to describe the structural/energetic features of MOFs. LCD is defined as the diameter of the largest sphere that can fit in the pore of MOF. The MaxER and MinER is the maximum and minimum value of variance explained by principal component analysis (PCA) for atom distribution in three directions of the unit cell, in which MaxER = MinER = 0.33 stands for the isotropic crystal. MPC/MNC is the most positive/negative charge of atoms in a unit cell. As for APC/ANC, the average positive/negative charge per unit volume was calculated. Furthermore, Aσ and Aε is the average σ/ε of an atom in the Lennard-Jones interaction. These descriptors were verified to possess significant correlations with the HCHO capture performance of MOFs.

4. Conclusions

In this work, we perform high-throughput computational screening of CoRE MOFs for HCHO capture with and without humidity conditions. In the dry air, working capacity and selectivity were adopted to evaluate the HCHO capture performance. It was found that small pore size (5–6 Å) and moderate heat of adsorption (40–50 kJ/mol) are favored for HCHO capture. Such high-performing structures probably have a 3D cage instead of a 2D channel with moderate charge and Lennard-jones parameters (0.2 ≤ APC ≤ 0.5, 1.5 ≤ Aε ≤ 3.0, and 0.12 ≤ Aσ ≤ 0.21) that benefit to the HCHO adsorption. Moreover, it was indicated that KH is the dominant factor to determine the HCHO capture performance, for which 10−2–101 mol/(kg·Pa) is preferred. The density plot of HCHO adsorption and adsorption isotherm verified that the top3 working capacity MOFs (LAVSUY, DUBWON, and PARMIG) are suitable for the removal of HCHO without H2O’s existence.
The SKH HCHO/H2O and SKH HCHO/(H2O + N2 + O2) was then obtained to screen MOFs under humidity condition. It was found that SKH HCHO/(H2O + N2 + O2) overestimates the influence of N2 and O2 ascribed to its high ratio in the air. It was suggested that MOFs with strong Coulombic interaction (high MPC, MNC) tends to have low SKH HCHO/H2O whereas the large Lennard-jones parameters (Aσ > 0.2 and Aε > 3) are required for MOFs exhibiting high SKH HCHO/H2O. Moreover, the adsorption isotherm of the top three structures (JAVTAC, WOJJOV, and ECAHAT) indicated that SKH HCHO/H2O can be recognized as a critical descriptor in low pressure, which all structures barely adsorb water. The simulation suggested that ECAHAT was a promising candidate for HCHO capture under 80% humidity conditions in 1 bar, 298 K, which have 0.64 mol/kg working capacity and high selectivity (reach 465).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms232213672/s1.

Author Contributions

W.L. methodology, writing and supervision; T.L. visualization; Y.L. data curation; W.W. writing—review and editing; S.L. supervision. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by the Fundamental Research Funds for the Central Universities (No. 21621039), Guang Zhou Basic and Applied Basic Research Foundation (No. 202201010433) and China Postdoctoral Science Foundation (2021M701413).

Conflicts of Interest

The authors declare that they have no conflict of interest.

References

  1. Kim, W.J.; Terada, N.; Nomura, T.; Takahashi, R.; Lee, S.D.; Park, J.H.; Konno, A. Effect of formaldehyde on the expression of adhesion molecules in nasal microvascular endothelial cells: The role of formaldehyde in the pathogenesis of sick building syndrome. Clin. Exp. Allergy 2002, 32, 287–295. [Google Scholar] [CrossRef] [PubMed]
  2. Wood, R.W.; Coleman, J.B. Behavioral evaluation of the irritant properties of formaldehyde. Toxicol. Appl. Pharmacol. 1995, 130, 67–72. [Google Scholar] [CrossRef]
  3. Gu, Z.-Y.; Wang, G.; Yan, X.-P. MOF-5 Metal-Organic Framework as Sorbent for in-Field Sampling and Preconcentration in Combination with Thermal Desorption GC/MS for Determination of Atmospheric Formaldehyde. Anal. Chem. 2010, 82, 1365–1370. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, D.; Xu, P.; Wang, H.; Zeng, G.; Huang, D.; Chen, M.; Lai, C.; Zhang, C.; Wan, J.; Xue, W. Strategies to improve metal organic frameworks photocatalyst’s performance for degradation of organic pollutants. Coord. Chem. Rev. 2018, 376, 449–466. [Google Scholar] [CrossRef]
  5. Szulejko, J.E.; Kim, K.-H.; Parise, J. Seeking the most powerful and practical real-world sorbents for gaseous benzene as a representative volatile organic compound based on performance metrics. Sep. Purif. Technol. 2019, 212, 980–985. [Google Scholar] [CrossRef]
  6. Lai, C.; Wang, Z.; Qin, L.; Fu, Y.; Li, B.; Zhang, M.; Liu, S.; Li, L.; Yi, H.; Liu, X.; et al. Metal-organic frameworks as burgeoning materials for the capture and sensing of indoor VOCs and radon gases. Coord. Chem. Rev. 2021, 427, 213565. [Google Scholar] [CrossRef]
  7. Dou, H.; Long, D.; Rao, X.; Zhang, Y.; Qin, Y.; Pan, F.; Wu, K. Photocatalytic Degradation Kinetics of Gaseous Formaldehyde Flow Using TiO2 Nanowires. ACS Sustain. Chem. Eng. 2019, 7, 4456–4465. [Google Scholar] [CrossRef]
  8. Yi, H.; Yan, M.; Huang, D.; Zeng, G.; Lai, C.; Li, M.; Huo, X.; Qin, L.; Liu, S.; Liu, X.; et al. Synergistic effect of artificial enzyme and 2D nano-structured Bi2WO6 for eco-friendly and efficient biomimetic photocatalysis. Appl. Catal. B Environ. 2019, 250, 52–62. [Google Scholar] [CrossRef]
  9. Yan, S.; Su, Y.; Deng, D.; Hu, J.; Lv, Y. Formaldehyde sensing based on high photoluminescence and strong oxidizing degradation of NH2-Fe(III)-nMOFs. Sens. Actuators B Chem. 2021, 333, 129140. [Google Scholar] [CrossRef]
  10. Zhang, S.; Zhuo, Y.; Ezugwu, C.I.; Wang, C.-c.; Li, C.; Liu, S. Synergetic Molecular Oxygen Activation and Catalytic Oxidation of Formaldehyde over Defective MIL-88B(Fe) Nanorods at Room Temperature. Environ. Sci. Technol. 2021, 55, 8341–8350. [Google Scholar] [CrossRef]
  11. Hu, S.-C.; Chen, Y.-C.; Lin, X.-Z.; Shiue, A.; Huang, P.-H.; Chen, Y.-C.; Chang, S.-M.; Tseng, C.-H.; Zhou, B. Characterization and adsorption capacity of potassium permanganate used to modify activated carbon filter media for indoor formaldehyde removal. Environ. Sci. Pollut. Res. 2018, 25, 28525–28545. [Google Scholar] [CrossRef] [PubMed]
  12. Yuan, X.; Deng, X.; Cai, C.; Shi, Z.; Liang, H.; Li, S.; Qiao, Z. Machine learning and high-throughput computational screening of hydrophobic metal–organic frameworks for capture of formaldehyde from air. Green Energy Environ. 2021, 6, 759–770. [Google Scholar] [CrossRef]
  13. Bellat, J.P.; Bezverkhyy, I.; Weber, G.; Royer, S.; Averlant, R.; Giraudon, J.M.; Lamonier, J.F. Capture of formaldehyde by adsorption on nanoporous materials. J. Hazard. Mater. 2015, 300, 711–717. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, K.J.; Miyawaki, J.; Shiratori, N.; Yoon, S.-H.; Jang, J. Toward an effective adsorbent for polar pollutants: Formaldehyde adsorption by activated carbon. J. Hazard. Mater. 2013, 260, 82–88. [Google Scholar] [CrossRef]
  15. Xu, Z.; Yu, J.; Xiao, W. Microemulsion-Assisted Preparation of a Mesoporous Ferrihydrite/SiO2 Composite for the Efficient Removal of Formaldehyde from Air. Chem. -A Eur. J. 2013, 19, 9592–9598. [Google Scholar] [CrossRef]
  16. Xu, Z.; Yu, J.; Liu, G.; Cheng, B.; Zhou, P.; Li, X. Microemulsion-assisted synthesis of hierarchical porous Ni(OH)2/SiO2 composites toward efficient removal of formaldehyde in air. Dalton Trans. 2013, 42, 10190–10197. [Google Scholar] [CrossRef]
  17. Xu, Z.; Yu, J.; Low, J.; Jaroniec, M. Microemulsion-Assisted Synthesis of Mesoporous Aluminum Oxyhydroxide Nanoflakes for Efficient Removal of Gaseous Formaldehyde. ACS Appl. Mater. Interfaces 2014, 6, 2111–2117. [Google Scholar] [CrossRef]
  18. Ewlad-Ahmed, A.M.; Morris, M.A.; Patwardhan, S.V.; Gibson, L.T. Removal of Formaldehyde from Air Using Functionalized Silica Supports. Environ. Sci. Technol. 2012, 46, 13354–13360. [Google Scholar] [CrossRef]
  19. Wen, Q.; Li, C.; Cai, Z.; Zhang, W.; Gao, H.; Chen, L.; Zeng, G.; Shu, X.; Zhao, Y. Study on activated carbon derived from sewage sludge for adsorption of gaseous formaldehyde. Bioresour. Technol. 2011, 102, 942–947. [Google Scholar] [CrossRef]
  20. Laszlo, K. Characterization and adsorption properties of polymer-based microporous carbons with different surface chemistry. Microporous Mesoporous Mater. 2005, 80, 205–211. [Google Scholar] [CrossRef]
  21. Ferey, G. Hybrid porous solids: Past, present, future. Chem. Soc. Rev. 2008, 37, 191–214. [Google Scholar] [CrossRef] [PubMed]
  22. Bobbitt, N.S.; Mendonca, M.L.; Howarth, A.J.; Islamoglu, T.; Hupp, J.T.; Farha, O.K.; Snurr, R.Q. Metal-organic frameworks for the removal of toxic industrial chemicals and chemical warfare agents. Chem. Soc. Rev. 2017, 46, 3357–3385. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, E.-X.; Yang, H.; Zhang, J. Zeolitic imidazolate framework as formaldehyde gas sensor. Inorg. Chem. 2014, 53, 5411–5413. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, D.; Li, Z.; Zhou, J.; Fang, H.; He, X.; Jena, P.; Zeng, J.-B.; Wang, W.-N. Simultaneous Detection and Removal of Formaldehyde at Room Temperature: Janus Au@ZnO@ZIF-8 Nanoparticles. Nano-Micro Lett. 2017, 10, 4. [Google Scholar] [CrossRef] [Green Version]
  25. Yoo, M.-J.; Lee, M.-H.; Szulejko, J.E.; Vikrant, K.; Kim, K.-H. A quantitation method for gaseous formaldehyde based on gas chromatography with metal–organic framework cold-trap sorbent as an effective alternative for HPLC-based standard protocol. Microchem. J. 2021, 160, 105624. [Google Scholar] [CrossRef]
  26. Wang, L.; Liang, X.-Y.; Chang, Z.-Y.; Ding, L.-S.; Zhang, S.; Li, B.-J. Effective Formaldehyde Capture by Green Cyclodextrin-Based Metal-Organic Framework. ACS Appl. Mater. Interfaces 2018, 10, 42–46. [Google Scholar] [CrossRef]
  27. Wang, Z.; Wang, W.; Jiang, D.; Zhang, L.; Zheng, Y. Diamine-appended metal-organic frameworks: Enhanced formaldehyde-vapor adsorption capacity, superior recyclability and water resistibility. Dalton. Trans. 2016, 45, 11306–11311. [Google Scholar] [CrossRef]
  28. Tran, T.Y.; Younis, S.A.; Heynderickx, P.M.; Kim, K.-H. Validation of two contrasting capturing mechanisms for gaseous formaldehyde between two different types of strong metal-organic framework adsorbents. J. Hazard. Mater. 2022, 424, 127459. [Google Scholar] [CrossRef]
  29. Boyd, P.G.; Chidambaram, A.; Garcia-Diez, E.; Ireland, C.P.; Daff, T.D.; Bounds, R.; Gladysiak, A.; Schouwink, P.; Moosavi, S.M.; Maroto-Valer, M.M.; et al. Data-driven design of metal-organic frameworks for wet flue gas CO2 capture. Nature 2019, 576, 253–256. [Google Scholar] [CrossRef] [Green Version]
  30. Bobbitt, N.S.; Snurr, R.Q. Molecular modelling and machine learning for high-throughput screening of metal-organic frameworks for hydrogen storage. Mol. Simul. 2019, 45, 1069–1081. [Google Scholar] [CrossRef]
  31. Budhathoki, S.; Ajayi, O.; Steckel, J.A.; Wilmer, C.E. High-throughput computational prediction of the cost of carbon capture using mixed matrix membranes. Energy Environ. Sci. 2019, 12, 1255–1264. [Google Scholar] [CrossRef]
  32. 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. [Google Scholar] [CrossRef]
  33. Lei, B.; 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. [Google Scholar]
  34. Daubert, T.E. Physical and Thermodynamic Properties of Pure Chemicals: Data Compilation; Design Institute for Physacal Property Data (DIPPR): New York, NY, USA, 1989. [Google Scholar]
  35. Wilmer, C.E.; Leaf, M.; Lee, C.Y.; Farha, O.K.; Hauser, B.G.; Hupp, J.T.; Snurr, R.Q. Large-scale Screening of Hypothetical Metal-organic Frameworks. Nat. Chem. 2012, 4, 83–89. [Google Scholar] [CrossRef] [PubMed]
  36. Banerjee, D.; Simon, C.M.; Plonka, A.M.; Motkuri, R.K.; Liu, J.; Chen, X.; Smit, B.; Parise, J.B.; Haranczyk, M.; Thallapally, P.K. Metal-organic framework with optimally selective xenon adsorption and separation. Nat. Commun. 2016, 7, ncomms11831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Ng, K.C.; Burhan, M.; Shahzad, M.W.; Ismail, A.B. A Universal Isotherm Model to Capture Adsorption Uptake and Energy Distribution of Porous Heterogeneous Surface. Sci. Rep. 2017, 7, 10634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Li, S.; Chung, Y.G.; Snurr, R.Q. High-Throughput Screening of Metal-Organic Frameworks for CO2 Capture in the Presence of Water. Langmuir 2016, 32, 10368–10376. [Google Scholar] [CrossRef]
  39. 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. [Google Scholar] [CrossRef]
  40. Manz, T.A.; Sholl, D.S. Improved Atoms-in-Molecule Charge Partitioning Functional for Simultaneously Reproducing the Electrostatic Potential and Chemical States in Periodic and Nonperiodic Materials. J. Chem. Theory Comput. 2012, 8, 2844–2867. [Google Scholar] [CrossRef]
  41. Nazarian, D.; Camp, J.S.; Sholl, D.S. A Comprehensive Set of High-Quality Point Charges for Simulations of Metal–Organic Frameworks. Chem. Mater. 2016, 28, 785–793. [Google Scholar] [CrossRef]
  42. Rappé, A.K.; Casewit, C.J.; Colwell, K.S.; Goddard, W.A., III; Skiff, W.M. UFF, a full periodic table force field for molecular mechanics and molecular dynamics simulations. J. Am. Chem. Soc. 1992, 114, 10024–10035. [Google Scholar] [CrossRef]
  43. Potoff, J.J.; Siepmann, J.I. Vapor-liquid equilibria of mixtures containing alkanes, carbon dioxide, and nitrogen. Aiche J. 2001, 47, 1676–1682. [Google Scholar] [CrossRef]
  44. Essmann, U.; Perera, L.; Berkowitz, M.L.; Darden, T.; Lee, H.; Pedersen, L.G. A smooth particle mesh Ewald method. J. Chem. Phys. 1995, 103, 8577–8593. [Google Scholar] [CrossRef] [Green Version]
  45. Ghosh, P.; Kim, K.C.; Snurr, R.Q. Modeling Water and Ammonia Adsorption in Hydrophobic Metal–Organic Frameworks: Single Components and Mixtures. J. Phys. Chem. C 2014, 118, 1102–1110. [Google Scholar] [CrossRef]
  46. Hantal, G.; Jedlovszky, P.; Hoang, P.N.M.; Picaud, S. Calculation of the adsorption isotherm of formaldehyde on ice by grand canonical Monte Carlo simulation. J. Phys. Chem. C 2007, 111, 14170–14178. [Google Scholar] [CrossRef]
Figure 1. Relationship between the largest cavity diameter (LCD) and (a) HCHO working capacity, (b) selectivity, colored by the heat of adsorption (Qst).
Figure 1. Relationship between the largest cavity diameter (LCD) and (a) HCHO working capacity, (b) selectivity, colored by the heat of adsorption (Qst).
Ijms 23 13672 g001
Figure 2. The relationship of formaldehyde capture performance and structural properties. (a) HCHO KH—working capacity relationship, colored by selectivity. (b) MaxER—working capacity relationship, colored by MinER.
Figure 2. The relationship of formaldehyde capture performance and structural properties. (a) HCHO KH—working capacity relationship, colored by selectivity. (b) MaxER—working capacity relationship, colored by MinER.
Ijms 23 13672 g002
Figure 3. Relationship between descriptors and formaldehyde capture performance, (a) correlation between working capacity (ΔW), and average positive charge (APC) of MOFs, colored by average negative charge (ANC). (b) correlation between selectivity (S) and average ε (Aε) of MOFs, colored by average of σ (Aσ).
Figure 3. Relationship between descriptors and formaldehyde capture performance, (a) correlation between working capacity (ΔW), and average positive charge (APC) of MOFs, colored by average negative charge (ANC). (b) correlation between selectivity (S) and average ε (Aε) of MOFs, colored by average of σ (Aσ).
Ijms 23 13672 g003
Figure 4. Formaldehyde, N2 and O2 adsorption isotherm of top-performing MOFs for a mixture component of HCHO/N2/O2 = 2/798/200 from GCMC simulation at 298 K ((a) for LAVSUY, (b) for DUBWON, (c) for PARMIG. The density distribution of formaldehyde adsorbates in (d) LAVSUY, (e) DUBWON and (f) PARMIG.
Figure 4. Formaldehyde, N2 and O2 adsorption isotherm of top-performing MOFs for a mixture component of HCHO/N2/O2 = 2/798/200 from GCMC simulation at 298 K ((a) for LAVSUY, (b) for DUBWON, (c) for PARMIG. The density distribution of formaldehyde adsorbates in (d) LAVSUY, (e) DUBWON and (f) PARMIG.
Ijms 23 13672 g004
Figure 5. Relationship between LCD and (a) SKH HCHO/H2O, (b) SKH HCHO/(H2O + N2 + O2), colored by the ASA.
Figure 5. Relationship between LCD and (a) SKH HCHO/H2O, (b) SKH HCHO/(H2O + N2 + O2), colored by the ASA.
Ijms 23 13672 g005
Figure 6. The relationship between chemical descriptors and Henry’s constant selectivity. (a) MPC and SKH HCHO/H2O, colored by the MNC; (b) Aε and SKH HCHO/(H2O + N2 + O2), colored by the Aσ.
Figure 6. The relationship between chemical descriptors and Henry’s constant selectivity. (a) MPC and SKH HCHO/H2O, colored by the MNC; (b) Aε and SKH HCHO/(H2O + N2 + O2), colored by the Aσ.
Ijms 23 13672 g006
Figure 7. HCHO, H2O, N2, and O2 adsorption isotherm of top-performing henry constant selectivity MOFs (a) JAVTAC, (b) WOJJOV, (c) ECAHAT, (d) DORDUK, (e) DOTTUC, (f) OHOMIH for a mixture component of HCHO/H2O/N2/O2 = 200/3280/77,216/19,304 from GCMC simulation at 298 K.
Figure 7. HCHO, H2O, N2, and O2 adsorption isotherm of top-performing henry constant selectivity MOFs (a) JAVTAC, (b) WOJJOV, (c) ECAHAT, (d) DORDUK, (e) DOTTUC, (f) OHOMIH for a mixture component of HCHO/H2O/N2/O2 = 200/3280/77,216/19,304 from GCMC simulation at 298 K.
Ijms 23 13672 g007aIjms 23 13672 g007b
Table 1. The largest cavity diameter (LCD), maximum explained ratio (MaxER), average positive charge (APC), average ε (Aε), Henry’s constant (KH), working capacity (ΔW) and selectivity (S) of the top 10 MOFs.
Table 1. The largest cavity diameter (LCD), maximum explained ratio (MaxER), average positive charge (APC), average ε (Aε), Henry’s constant (KH), working capacity (ΔW) and selectivity (S) of the top 10 MOFs.
REFCODELCD
Å
MaxERAPC
e

kcal/mol
KH
mol/(kg·Pa)
ΔW
mol/kg
S
LAVSUY6.620.430.500.161.18 × 10−24.012722
DUBWON5.200.380.480.144.19 × 10−23.988189
PARMIG4.710.370.240.231.72 × 10−23.934044
SEHTAB5.170.470.400.133.16 × 10−23.823157
DEYJIC4.950.560.330.174.98 × 10−23.687689
ADIQEL4.250.370.180.221.01 × 10−23.601453
LIFWOO4.980.380.160.212.76 × 10−23.343486
DEFKUU5.420.440.590.187.37 × 10−33.3112,015
NABMUA016.100.430.440.181.17 × 10−13.152180
LOBHAM6.510.390.270.172.68 × 10−13.094160
Table 2. The LCD, MPC, MNC, and KH of HCHO, H2O, N2, and O2 for the top 10 MOFs in SKH HCHO/H2O.
Table 2. The LCD, MPC, MNC, and KH of HCHO, H2O, N2, and O2 for the top 10 MOFs in SKH HCHO/H2O.
REFCODELCD
Å
MPC
e
MNC
e
KH HCHO
mol/(kg·Pa)
KH H2O
mol/(kg·Pa)
KH N2
mol/(kg·Pa)
KH O2
mol/(kg·Pa)
SKH
HCHO/H2O
JAVTAC5.082.44−1.144.70 × 10−16.85 × 10−58.96 × 10−58.71 × 10−5418.76
WOJJOV7.811.72−0.685.16 × 10−21.62 × 10−52.29 × 10−52.32 × 10−5194.11
ECAHAT12.440.88−0.612.08 × 10−28.75 × 10−61.13 × 10−51.38 × 10−5144.74
PUQYAC5.331.03−0.622.10 × 10−31.88 × 10−68.87 × 10−69.35 × 10−668.04
LIDZUV4.491.6−0.781.02 × 10−21.10 × 10−52.10 × 10−52.23 × 10−556.81
ZERQOE4.241.59−0.784.56 × 10−35.39 × 10−61.50 × 10−51.93 × 10−551.59
KAXQOR4.231.59−0.784.07 × 10−35.22 × 10−61.41 × 10−51.83 × 10−547.56
IXISOX5.570.24−0.363.84 × 10−25.12 × 10−59.73 × 10−61.04 × 10−545.73
PARMIG4.710.76−0.581.72 × 10−22.35 × 10−54.37 × 10−55.31 × 10−544.6
GUPBEZ7.290.1−0.313.72 × 10−35.64 × 10−62.01 × 10−62.26 × 10−640.24
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, W.; Liang, T.; Lin, Y.; Wu, W.; Li, S. In Silico Screening of Metal-Organic Frameworks for Formaldehyde Capture with and without Humidity by Molecular Simulation. Int. J. Mol. Sci. 2022, 23, 13672. https://doi.org/10.3390/ijms232213672

AMA Style

Li W, Liang T, Lin Y, Wu W, Li S. In Silico Screening of Metal-Organic Frameworks for Formaldehyde Capture with and without Humidity by Molecular Simulation. International Journal of Molecular Sciences. 2022; 23(22):13672. https://doi.org/10.3390/ijms232213672

Chicago/Turabian Style

Li, Wei, Tiangui Liang, Yuanchuang Lin, Weixiong Wu, and Song Li. 2022. "In Silico Screening of Metal-Organic Frameworks for Formaldehyde Capture with and without Humidity by Molecular Simulation" International Journal of Molecular Sciences 23, no. 22: 13672. https://doi.org/10.3390/ijms232213672

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