Next Article in Journal
Integrating Pressure-Driven Membrane Separation Processes to Improve Eco-Efficiency in Cheese Manufacture: A Preliminary Case Study
Next Article in Special Issue
The Impact of Various Natural Gas Contaminant Exposures on CO2/CH4 Separation by a Polyimide Membrane
Previous Article in Journal
A Review of CFD Modelling and Performance Metrics for Osmotic Membrane Processes
Previous Article in Special Issue
Synthesis of Cis-Cisoid or Cis-Transoid Poly(Phenyl-Acetylene)s Having One or Two Carbamate Groups as Oxygen Permeation Membrane Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Selective Hydrogen Permeation through Graphdiyne Membrane: A Theoretical Study

1
Analytical and Testing Center, Anhui University of Science and Technology, Huainan 232001, China
2
State Key Laboratory of Materials-Oriented Chemical Engineering, College of Chemical Engineering, Nanjing Tech University, 30 Puzhu Road (S), Nanjing 211816, China
*
Authors to whom correspondence should be addressed.
Membranes 2020, 10(10), 286; https://doi.org/10.3390/membranes10100286
Submission received: 29 September 2020 / Revised: 9 October 2020 / Accepted: 13 October 2020 / Published: 15 October 2020
(This article belongs to the Special Issue Polymer Membranes for Gas Separation)

Abstract

:
Graphdiyne (GDY), with uniform pores and atomic thickness, is attracting widespread attention for application in H2 separation in recent years. However, the challenge lies in the rational design of GDYs for fast and selective H2 permeation. By MD and DFT calculations, several flexible GDYs were constructed to investigate the permeation properties of four pure gas (H2, N2, CO2, and CH4) and three equimolar binary mixtures (H2/N2, H2/CO2, and H2/CH4) in this study. When the pore size is smaller than 2.1 Å, the GDYs acted as an exceptional filter for H2 with an approximately infinite H2 selectivity. Beyond the size-sieving effect, in the separation process of binary mixtures, the blocking effect arising from the strong gas–membrane interaction was proven to greatly impede H2 permeation. After understanding the mechanism, the H2 permeance of the mixtures of H2/CO2 was further increased to 2.84 × 105 GPU by reducing the blocking effect with the addition of a tiny amount of surface charges, without sacrificing the selectivity. This theoretical study provides an additional atomic understanding of H2 permeation crossing GDYs, indicating that the GDY membrane could be a potential candidate for H2 purification.

Graphical Abstract

1. Introduction

As an attractive alternative fuel source, hydrogen (H2) could eliminate the use of polluting fossil fuels in industry and transport in the future [1]. So, the H2 energy is critical to reduce global greenhouse gases and promote sustainable development because of its natural abundance and high efficiency of combustion [2,3]. Today, in many industrial and drilling streams, H2 is mainly produced from natural gas, hydro-electrolysis, and the combustion of hydrides [4,5]. However, these processes release several million tonnes of by-products (such as carbon dioxide, nitrogen, and methane) per year [6]. How to separate the target product of H2 from the undesirable species is crucial to improve production efficiency and reduce the cost. Traditional separation techniques, such as pressure swing adsorption and cryogenic separation, would consume a considerable amount of energy to collect H2 [7]. With the improved performances and lower operating conditions, the advanced membrane-based separation is seen as an alternative to significantly improve energy efficiency [8,9]. At room temperature and low transmembrane pressure, it is more instructive for experiments to develop membranes with both good H2 permeance and selectivity. [10].
Recently, two-dimensional (2D) carbon-based membranes have sparked global attention due to their atomic thickness [11]. Among various membranes, graphene-based materials are, assuredly, one of the powerful candidates for membrane separation according to our previous experimental [12,13] and simulation studies [14,15]. According to the separation mechanism of the size sieving effect, controllable pores in the 2D-material membrane are imperative for gas separation. However, it is not cost-effective to drill uniform pores in graphene, and this process may introduce non-selective defects in the membrane [16]. Alternatively, newly emerged graphdiyne (GDY) allotropes that formed by the periodic combination of sp and sp2 carbon atoms are proposed as alternative options for gas purification [16], because they process the well-distributed pores as well as the atomic membrane thickness [17]. Several studies have been reported to explore the gas permeation property through GDYs in recent years [18,19,20]. It is the comparable pores in GDYs that contribute to the outstanding gas separation properties, as Smith et al. first reported that the excellent H2 separation performance was ascribed to the small triangular pores in the GDYs [21]. The later reports also showed that the various sized pores in diverse GDYs could be employed to distinguish the different sized molecules, such as He [18], O2 [22], and CO2 [23,24], concluding that the size sieving effect dominated the transport mechanism of GDYs for gas separation.
Although the pore size is an important factor for gas separation membranes, other factors, particularly the surface properties, should not be neglected. Sang et al. [25] and Smith et al. [26] found that the functionalized surface of GDYs shown a better separation performance than that of the pristine one. Moreover, the adsorption phenomenon in GDYs also affected the permeation of gases, such as H2S [27] and CO2 [24,28]. It means, beyond the well-understood size sieving effect, there could be a more comprehensive mechanism to better describe gas transport through GDYs [19]. However, the main challenge is how such a mechanism determines the gas permeation properties, especially for selective H2 permeation through GDYs, and how it contributes to further improvement of the H2 permeance and selectivity of GDYs. In addition, most previous works focused on the first-principles calculations density functional theory (DFT) to study the selectivity of H2 over CO2, CO, N2, and CH4 [25,26,29]. A few studies calculated the H2 permeance by performing molecular dynamic (MD) simulations, which however were based on a rigid framework of GDYs [20,23,30] and might be too idealized for actual GDYs for gas separation. Moreover, it is unclear what the ultimate size of nanopores in GDYs is allowed to transport H2 molecules. Therefore, it is necessary to further understand the underlying separation mechanism of H2 purification through a carefully designed flexible GDY membrane that possesses both high H2 permeance and selectivity and find the ultimate diameter in GDYs for H2 permeation.
In this work, a series of 2D GDYs are computationally constructed to examine the permeation of four pure gases (H2, N2, CO2 and CH4) and their equimolar binary mixtures (H2/N2, H2/CO2 and H2/CH4). Assuming that the separation performance of H2 is affected by pore structures and surface charges, we systematically investigated these two effects by both MD and DFT calculations. Following the introduction, the atomic models of GDYs, as well as the separation systems, are illustrated in Section 2. In Section 3, the H2 permeance is calculated by the time evolution of the permeated molecules according to the MD results, and the ideal selectivity is evaluated by DFT calculations. The separation mechanism of H2 from binary mixtures is revealed by analyzing the diffusion coefficient, density contour, and energy barrier of the permeation. After that, the H2 permeance of the mixture of H2/CO2 is further improved by reducing the blocking effect. Finally, the concluding remarks are summarized in Section 4.

2. Models and Methods

The GDYs were constructed in the Material studio [31] and then subjected to geometry optimization in the Forcite module with 5000 iterations. As presented in Figure 1a–d, the dimensions of the membranes were 7.75 × 7.55 nm2 in this setup, and the investigated pore diameters were varied from 1.5 to 2.5 Å, which were measured according to the formula: D =   2 A / π by inserting a van der Waals sphere, where A is the open pore area, as depicted in Figure 1e–h. We noted that the membrane in Figure 1b coded as GDY_1.5Å_p7% has the porosity of 7%, which is larger than that of 3% in Figure 1a (GDY_1.5Å_p3%), although both of these two membranes have a similar pore diameter of 1.5 Å. The above-mentioned porosity was calculated by the formula p = A p o r e A m e m , where A p o r e and A m e m denote to the areas of the totally unoccupied regions and the membrane surface, respectively.
Based on the optimized membranes, MD calculations were performed to simulate the gas separation. Figure 2 shows the simulation system, where two chambers were isolated by the GDY membrane. The left chamber was a gas reservoir, comprising 2000 molecules of pure gases (or binary mixtures with the mixing ratio of 1000:1000, in volume ratio). The right one treated as a vacuum is the permeate side, which is the most common setting in MD simulations for collecting the permeated molecules. [20,23,30] At both ends of each chamber, rigid graphene was placed to prevent molecules from roaming between the periodic boxes. Considering that carbon-based membranes usually have good flexibility in experiments [32], all pores in GDYs were treated as flexible in our present work so that the atoms around pores could perform small displacements. Thus, the aperture would be enlarged to allow the passage of H2 even though the molecular size is larger than the pore size. On the contrary, the atoms on the edge of membranes were imposed with a position restriction to ensure the uniform lattice of GDYs and decrease the impact on gas permeation [33]. Furthermore, no collapse appeared in our system, suggesting good stability of the flexible structures. The framework of GDYs was described by the all-atom optimized potential. The four gases were represented by Lennard–Jones (LJ) and electrostatic potentials originating from our previous work [14]. Between dissimilar atoms, the interactions were assembled by the Lorentz–Berthelot combination rule [34].
In this study, MD simulations were all carried out using the GROMACS package (version 4.5.5) [35]. A static minimization with the steepest descent method was first carried out to remove the unreasonable contacts among each atom. Then, the separation system was pre-equilibrated in the isobaric–isothermal ensemble (constant temperature and constant pressure ensemble, named as NPT) for 2 ns only in the z-direction. Following this, another 20 ns MD simulations were performed in the canonical ensemble (constant temperature and constant volume ensemble, named as NVT) for data collection and further analysis, and the trajectories were stored every 1 ps. During the simulations, the classical equations of motion were integrated with a time step of 1 fs by using the leapfrog algorithm. The temperature was maintained at 300 K by coupling with the velocity rescaling [36] thermostats. The long-range electrostatic interactions were computed by using the method of particle mesh Ewald [37]. While the short-range van der Waals interactions were truncated at a cut-off distance of 1.2 nm, the periodic boundary conditions were implemented in all three directions. With the partial pressure gradient as the driving force [38], the gases would permeate through the GDYs. For a better understanding of the separation process, an animation is provided in the Supplementary Materials (Video S1).

3. Results and Discussion

3.1. Gas Permeation Behavior

The relationship between gas flux ( J , mol·m−2·s−1) and permeance ( S , mol·m−2·s−1·Pa−1) was described as Equation (1):
J = d N A m e m N A d t = P S
where N refers to the number of permeated molecules, N A is the Avogadro constant, and P is the transmembrane gas partial pressure estimated by Equation (2):
P = ( N 0 N a d N ) k T V l N k T V r
in which N a d refers to the number of gases adsorbed on the membrane surface, V l and V r are the volumes of the left and right chambers, respectively, and k represents the Boltzmann constant. Therefore, the time evolution of the number of permeated molecules is integrated as Equation (3), where R is the gas constant,
N = ( N 0 N a d ) L r ( L l + L r ) ( 1 e RT S ( L l + L r ) L l L r t ) = 2 ( N 0 N a d ) 3 ( 1 e 467.7 S t ) .
As presented in Figure 3a, the number of permeated H2 is increased exponentially with the simulation time, which agrees well with the above mathematical analysis. The permeance of pure H2 is shown in Figure 3b. Obviously, it is remarkably enhanced with the increase of porosity and pore size. With the smallest pore of 1.5 Å, the porosity of 0.03 and 0.07 have the H2 permeance of 1.34   ×   105 and 2.55   ×   105 GPU (1 GPU = 3.35   ×   10−10 mol·m−2·s−1·Pa−1), respectively. The permeation of other gases (i.e., CO2, N2 and CH4) through the GDYs was also simulated. It was shown that only the biggest pore with 2.5 Å allowed the passage of N2, CO2, and CH4, as presented in Figure 3c. The incompatibility of the kinetic diameter of gases (i.e., H2, 2.89 Å; N2, 3.64 Å; CO2, 3.30 Å; CH4, 3.8 Å) [39], and the pore size is ascribed to the flexible structures, implying that the membrane of GDY_2.5Å is not suitable for H2 purification. Moreover, the comparable pore with 2.1 Å diameter can not only completely block the other three gases but also process the good H2 permeance of 5.94   ×   105 GPU, which implies that the selectivities of H2 over CO2, N2, and CH4 can be extremely high in the membrane of GDY_2.1Å. We noted that the molecules of N2, CO2, and CH4 can not be detected in the permeation side as long as the pore diameter was smaller than 2.1 Å, indicating the infinitely low permeance of N2, CO2, and CH4.
To evaluate the ideal selectivities of H2 over other three gases in GDY_2.1 Å, the DFT calculations were performed to calculate permeation barriers as per our previous study [15]. Figure 4a illustrates the minimum energy pathway (MEP) of four gases crossing the membrane. The inset configurations are the energetically stable states of CO2 permeation at different locations. By searching the saddle point in MEPs, the energy barriers of permeation can be calculated. Evidently, the permeation barrier of H2 crossing the flexible GDY_2.1 Å is drastically reduced to 1.55 kJ/mol (Figure 4a, inset), which results in extraordinary H2 permeance. On the contrary, the permeation barriers of CO2, N2, and CH4 are all extremely high. In Figure 4b, the temperature-dependent H2 selectivity has an inverse correlation with the temperature according to the Arrhenius Equation (4), whereas Pi is the permeation rate, A is the permeation prefactor that can be assumed as 1011 s−1 [38], and T is the temperature. For H2/CO2 and H2/N2, the ideal selectivities of H2 can be up to 1017 at 300 K, which is the same order with the modified GDYs [25]. It remains very high (>108) even at 600 K, further suggesting the extraordinary H2 separation performance through GDY_2.1 Å.
S i / j = P i P j = A i e x p ( E b a r r i e r ,     i R T ) A j e x p ( E b a r r i e r ,     j R T )

3.2. Transport Mechanism: Blocking Effect

For binary gas mixtures, the MD calculations were also performed to simulate the separation process. To visualize H2 purification properties through the membrane of GDY_2.1Å, the equilibrium configurations of final frames are presented in Figure 5a–c. As seen, the GDY_2.1Å membrane acts as an effective filter for H2 separation while completely blocking the passage of CO2, N2, and CH4, so that the H2 is largely gathering in the vacuum chamber. The corresponding H2 permeance of the three binary mixtures is presented in Figure 5d. Interestingly, the mixture of H2/N2 exhibits the highest H2 permeance of 4.71 × 105 GPU, which is a little higher than that of H2/CH4 and almost twice as many as the mixture of H2/CO2. The primary reason is that on the membrane surface, the strongly interacting gas (CO2) preferentially adsorbs, blocking the transport pores of H2 molecules, thus resulting in a relatively low H2 permeance.
Further analyses are carried out to understand this blocking effect. The radial distribution function (RDF) was calculated to present the affinity of GDY_2.1Å to four gases by using Equation (5):
g i j ( r ) = N i j ( r ,   r + r ) V 4 π r 2 r N i N j
where Nij (r,r + Δr) is the number of species j around i within a shell from r to r + Δr, r is the distance between two species, and Ni and Nj refer to the numbers of atom types i and j, respectively. As presented in Figure 6a, there is an increasing trend of the gas–membrane interaction as H2   <   N2   <   CH4   <   CO2. The weakest interaction together with the smallest molecular size endows H2 with exceptional permeance. Meanwhile, the strong interaction promotes the gases, particularly CO2 to preferentially adsorb on the membrane surface. To offer more intuitionistic information, the density contours of the distribution of N2, CH4, and CO2 were plotted on the GDY_2.1Å surface. The general distribution behavior of the three gases is similar as shown in Figure 6b–d, where the pores that are largely clogged by these gases are larger than the pore diameters. Nevertheless, the intensity of gas accumulation is quite different. As seen in Figure 6d, the pores are the most clogged, with the number of adsorbed CO2 molecules exceeding 6.5 Nw/uc in every pore. Similar to the affinity analysis in Figure 6a, the extent of the blocking effect follows the increasing trend of N2   <   CH4   <   CO2, decreasing the placeholders of H2 on the membrane surface in Figure 7a.
According to the first peak in Figure 6a, the mean square displacement (MSD) of H2 molecules was analyzed within 0.5 nm of the membrane surface (Figure 7b, inset) from
M S D ( t )   = 1 N i = 1 N ( r i ( t ) 2 r i ( t 0 ) 2 )
D =   1 6 l i m t d M S D ( t ) d t
where r i ( t ) 2 r i ( t 0 ) 2 refers to the distance traveled by the atom (i) over the time interval of t t 0 . The diffusion coefficient (D) is determined by the linear slope of the MSD curve with Equation (7). The diffusion behavior in this confined region cannot sustain a normal movement with a large time interval, and it is sufficient to calculate the diffusion coefficient within 20 ps. The greater placeholders of H2, ascribing to the lower blocking effect of N2, promote the faster movement of H2, as presented in Figure 7b. H2 in three binary mixtures of H2/N2, H2/CH4, and H2/CO2 exhibits the diffusion coefficients of 0.57 × 10−2, 0.44 × 10−2, and 0.33 × 10−2 cm2/s, respectively. That is, the H2 permeation is indeed impeded by the preferentially adsorbed CO2 molecules. In other words, the GDY_2.1Å can separate H2 faster and more selectively from such a kind of binary mixture that the other species has a weak interaction toward GDY surfaces, such as H2/N2.

3.3. Surface Charge Effect

Two important mechanisms dominated the gas separation. The first one is the size sieving effect, which favors the permeation of small-sized molecules of H2. While beyond the size sieving effect, in the separation process of binary mixtures, the blocking effect greatly impedes H2 permeation. The above understanding of the separation mechanism is beneficial to further guide the improvement of H2 permeance particularly for the mixture of H2/CO2 without sacrificing its selectivity. An effective way proposed here is to reduce the most severe blocking effect of CO2 by surface charge modification. As depicted in the inset figure in Figure 8a, the positive and negative charges are uniformly imposed on the network of GDY_2.1Å from ±0.00001 to ±0.35 e/atom, whereas the net charge of the whole membrane is zero. The CO2 molecules still cannot cross the membrane regardless of surface charges. Compared to the pristine GDY membrane, the increased H2 permeance of the binary mixture of H2/CO2 was observed when surface charges are lower than ±0.1 e/atom in Figure 8b. It can be up to 2.84 × 105 GPU by imposing a tiny amount of surface charge of ±0.035–0.050 e/atom. This exceptionally high permeance is several orders of magnitude greater than the existed experiments [40,41,42], which is ascribed to the ultimate pore size and atomic thickness of GDYs. The following decreasing trend in Figure 8b was ascribed to the surface overcharge, where the generated strong electrostatic repels not only CO2 but also H2 from approaching. Thus, achieving the perfect balance between these two mechanisms, a tiny amount of surface charges is qualified to maximize the H2 permeance by reducing the blocking effect of strong interlaced molecules of CO2, meanwhile ensuring the placeholders of small-sized molecules of H2 as well.

4. Conclusions

In summary, a multiscale study combining MD and DFT calculations was performed to investigate the gas permeation through carefully designed flexible GDY membranes with different pore structures and surface charges. Four single gases and three equimolar binary mixtures (H2/other gas) were simulated to study the selective gas permeation through GDYs. Approximately infinite selectivities of H2 over N2, CO2, and CH4 in GDY_2.1Å membranes were demonstrated by DFT calculations. The underlying mechanism indicated that the blocking effect impeded H2 permeation and the GDY_2.1Å is prone to separate H2 from such binary mixtures in which the other species has a weak gas–membrane interaction. Moreover, by imposing a tiny amount of surface charges, the H2 permeance of the binary H2/CO2 mixture was further enhanced up to 2.84 × 105 GPU without sacrificing the selectivity. These excellent transport properties make the GDYs a promising candidate for efficient H2 purification. Although the present work is a theoretical study, it is believable that the GDYs can be elaborately designed for realistic separations with the improvement of a well-controlled synthesis strategy.

Supplementary Materials

The following are available online at https://www.mdpi.com/2077-0375/10/10/286/s1, Video S1: The separation process of the binary mixture of H2/CO2 through GDY_2.1Å.

Author Contributions

Conceptualization, Q.L. and G.L.; Data curation, Q.L.; Formal analysis, Q.L.; Funding acquisition, G.L.; Investigation, Q.L.; Methodology, Q.L.; Resources, G.L.; Writing—original draft, Q.L.; Writing—review & editing, L.C. and G.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (grant numbers 21922805, 21776125). We are grateful to the High-Performance Computing Center of Nanjing Tech University for supporting the computational resources.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Corredor, J.; Perez-Pena, E.; Rivero, M.J.; Ortiz, I. Performance of rGO/TiO2 photocatalytic membranes for hydrogen production. Membranes 2020, 10, 218. [Google Scholar] [CrossRef] [PubMed]
  2. Staffell, I.; Scamman, D.; Velazquez Abad, A.; Balcombe, P.; Dodds, P.E.; Ekins, P.; Shah, N.; Ward, K.R. The role of hydrogen and fuel cells in the global energy system. Energy Environ. Sci. 2019, 12, 463–491. [Google Scholar] [CrossRef] [Green Version]
  3. Parvasi, P.; Mohammad Jokar, S.; Basile, A.; Iulianelli, A. An on-board pure H2 supply system based on a membrane reactor for a fuel cell vehicle: A theoretical study. Membranes 2020, 10, 159. [Google Scholar] [CrossRef]
  4. Zhang, Y.; Xu, P.; Liang, S.; Liu, B.; Shuai, Y.; Li, B. Exergy analysis of hydrogen production from steam gasification of biomass: A review. Int. J. Hydrog. Energy 2019, 44, 14290–14302. [Google Scholar] [CrossRef]
  5. Franchi, G.; Capocelli, M.; De Falco, M.; Piemonte, V.; Barba, D. Hydrogen production via steam reforming: A critical analysis of MR and RMM technologies. Membranes 2020, 10, 10. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, Y.; Lee, W.H.; Seong, J.G.; Bae, J.Y.; Zhuang, Y.; Feng, S.; Wan, Y.; Lee, Y.M. Alicyclic segments upgrade hydrogen separation performance of intrinsically microporous polyimide membranes. J. Membr. Sci. 2020, 611, 118363. [Google Scholar] [CrossRef]
  7. Ockwig, N.W.; Nenoff, T.M. Membranes for hydrogen separation. Chem. Rev. 2007, 107, 4078–4110. [Google Scholar] [CrossRef]
  8. Casadei, R.; Giacinti Baschetti, M.; Yoo, M.J.; Park, H.B.; Giorgini, L. Pebax@2533/graphene oxide nanocomposite membranes for carbon capture. Membranes 2020, 10, 188. [Google Scholar] [CrossRef]
  9. Castel, C.; Favre, E. Membrane separations and energy efficiency. J. Membr. Sci. 2018, 548, 345–357. [Google Scholar] [CrossRef]
  10. Richard, W. Baker. Overview of membrane science and technology. In Membrane Technology and Applications; John Wiley & Sons, Ltd.: Chichester, UK, 2012; pp. 1–14. [Google Scholar]
  11. Wang, S.; Yang, L.; He, G.; Shi, B.; Li, Y.; Wu, H.; Zhang, R.; Nunes, S.; Jiang, Z. Two-dimensional nanochannel membranes for molecular and ionic separations. Chem. Soc. Rev. 2020, 49, 1071–1089. [Google Scholar] [CrossRef]
  12. Liu, G.; Jin, W.; Xu, N. Graphene-based membranes. Chem. Soc. Rev. 2015, 44, 5016–5030. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, G.; Jin, W.; Xu, N. Two-Dimensional-Material Membranes: A new family of high-performance separation membranes. Angew. Chem. 2016, 55, 13384–13397. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, Q.; Gupta, K.M.; Xu, Q.; Liu, G.; Jin, W. Gas permeation through double-layer graphene oxide membranes: The role of interlayer distance and pore offset. Sep. Purif. Technol. 2019, 209, 419–425. [Google Scholar] [CrossRef]
  15. Liu, Q.; Liu, Y.; Liu, G. Simulation of cations separation through charged porous graphene membrane. Chem. Phys. Lett. 2020, 753, 137606. [Google Scholar] [CrossRef]
  16. James, A.; John, C.; Owais, C.; Myakala, S.N.; Chandra Shekar, S.; Choudhuri, J.R.; Swathi, R.S. Graphynes: Indispensable nanoporous architectures in carbon flatland. RSC Adv. 2018, 8, 22998–23018. [Google Scholar] [CrossRef] [Green Version]
  17. Gao, X.; Liu, H.; Wang, D.; Zhang, J. Graphdiyne: Synthesis, properties, and applications. Chem. Soc. Rev. 2019, 48, 908–936. [Google Scholar] [CrossRef]
  18. Rezaee, P.; Naeij, H.R. Graphenylene–1 membrane: An excellent candidate for hydrogen purification and helium separation. Carbon 2020, 157, 779–787. [Google Scholar] [CrossRef] [Green Version]
  19. Qiu, H.; Xue, M.; Shen, C.; Zhang, Z.; Guo, W. Graphynes for water desalination and gas separation. Adv. Mater 2019, 31, e1803772. [Google Scholar] [CrossRef]
  20. Mahdizadeh, S.J.; Goharshadi, E.K. Multicomponent gas separation and purification using advanced 2D carbonaceous nanomaterials. RSC Adv. 2020, 10, 24255–24264. [Google Scholar] [CrossRef]
  21. Jiao, Y.; Du, A.; Hankel, M.; Zhu, Z.; Rudolph, V.; Smith, S.C. Graphdiyne: A versatile nanomaterial for electronics and hydrogen purification. Chem. Commun. 2011, 47, 11843–11845. [Google Scholar] [CrossRef]
  22. Meng, Z.; Zhang, X.; Zhang, Y.; Gao, H.; Wang, Y.; Shi, Q.; Rao, D.; Liu, Y.; Deng, K.; Lu, R. Graphdiyne as a high-efficiency membrane for separating oxygen from harmful gases: A first-principles study. ACS Appl. Mater. Interfaces 2016, 8, 28166–28170. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, L.; Sang, P.; Guo, S.; Liu, X.; Li, J.; Zhu, H.; Guo, W. Promising monolayer membranes for CO2/N2/CH4 separation: Graphdiynes modified respectively with hydrogen, fluorine, and oxygen atoms. Appl. Surf. Sci. 2017, 405, 455–464. [Google Scholar] [CrossRef]
  24. Apriliyanto, Y.B.; Lago, N.F.; Lombardi, A.; Evangelisti, S.; Bartolomei, M.; Leininger, T.; Pirani, F. Nanostructure selectivity for molecular adsorption and separation: The case of graphyne layers. J. Phys. Chem. C 2018, 122, 16195–16208. [Google Scholar] [CrossRef]
  25. Sang, P.; Zhao, L.; Xu, J.; Shi, Z.; Guo, S.; Yu, Y.; Zhu, H.; Yan, Z.; Guo, W. Excellent membranes for hydrogen purification: Dumbbell-shaped porous γ-graphynes. Int. J. Hydrog. Energy 2017, 42, 5168–5176. [Google Scholar] [CrossRef]
  26. Jiao, Y.; Du, A.; Smith, S.C.; Zhu, Z.; Qiao, S.Z. H2 purification by functionalized graphdiyne–role of nitrogen doping. J. Mater. Chem. A 2015, 3, 6767–6771. [Google Scholar] [CrossRef]
  27. Lei, G.; Liu, C.; Li, Q.; Xu, X. Graphyne nanostructure as a potential adsorbent for separation of H2S/CH4 mixture: Combining grand canonical Monte Carlo simulations with ideal adsorbed solution theory. Fuel 2016, 182, 210–219. [Google Scholar] [CrossRef]
  28. Bartolomei, M.; Giorgi, G. A nnovel nanoporous graphite based on graphynes: First-principles structure and carbon dioxide preferential physisorption. ACS Appl. Mater. Interfaces 2016, 8, 27996–28003. [Google Scholar] [CrossRef] [Green Version]
  29. Zhang, H.; He, X.; Zhao, M.; Zhang, M.; Zhao, L.; Feng, X.; Luo, Y. Tunable hydrogen separation in sp-sp2 hybridized carbon membranes: A first-principles prediction. J. Phys. Chem. C 2012, 116, 16634–16638. [Google Scholar] [CrossRef]
  30. Bartolomei, M.; Carmona-Novillo, E.; Hernández, M.I.; Campos-Martínez, J.; Pirani, F.; Giorgi, G. Graphdiyne pores: “Ad Hoc” openings for helium separation applications. J. Phys. Chem. C 2014, 118, 29966–29972. [Google Scholar] [CrossRef] [Green Version]
  31. Delley, B. An all-electron numerical method for solving the local density functional for polyatomic molecules. J. Chem. Phys. 1990, 92, 508–517. [Google Scholar] [CrossRef]
  32. Lee, S.-M.; Kim, J.-H.; Ahn, J.-H. Graphene as a flexible electronic material: Mechanical limitations by defect formation and efforts to overcome. Mater. Today 2015, 18, 336–344. [Google Scholar] [CrossRef]
  33. Zhou, K.; Xu, Z. Renormalization of ionic solvation shells in nanochannels. ACS Appl. Mater. Interfaces 2018, 10, 27801–27809. [Google Scholar] [CrossRef]
  34. Wennberg, C.L.; Murtola, T.; Páll, S.; Abraham, M.J.; Hess, B.; Lindahl, E. Direct-space corrections enable fast and accurate Lorentz–Berthelot combination rule Lennard-Jones lattice summation. J. Chem. Theory Comput. 2015, 11, 5737–5746. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS 4:  Algorithms for highly efficient, load-balanced, and scalable molecular simulation. J. Chem. Theory Comput. 2008, 4, 435–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling through velocity rescaling. J. Chem. Phys. 2007, 126, 014101. [Google Scholar] [CrossRef] [Green Version]
  37. 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]
  38. Xu, J.; Zhou, S.; Sang, P.; Li, J.; Zhao, L. Inorganic graphenylene as a promising novel boron nitrogen membrane for hydrogen purification: A computational study. J. Mater. Sci. 2017, 52, 10285–10293. [Google Scholar] [CrossRef]
  39. Van Reis, R.; Zydney, A. Bioprocess membrane technology. J. Membr. Sci. 2007, 297, 16–50. [Google Scholar] [CrossRef]
  40. Omidvar, M.; Nguyen, H.; Liang, H.; Doherty, C.M.; Hill, A.J.; Stafford, C.M.; Feng, X.; Swihart, M.T.; Lin, H. Unexpectedly strong size-sieving ability in carbonized polybenzimidazole for membrane H2/CO2 separation. ACS Appl. Mater. Interfaces 2019, 11, 47365–47372. [Google Scholar] [CrossRef]
  41. Sun, Y.; Song, C.; Guo, X.; Liu, Y. Concurrent Manipulation of out-of-plane and regional in-plane orientations of NH2-UiO-66 membranes with significantly reduced anisotropic grain boundary and superior H2/CO2 separation performance. ACS Appl. Mater. Interfaces 2020, 12, 4494–4500. [Google Scholar] [CrossRef]
  42. Robeson, L.M. The upper bound revisited. J. Membr. Sci. 2008, 320, 390–400. [Google Scholar] [CrossRef]
Figure 1. Membrane models. Atomic structures of graphdiyne (GDY) membranes with different pore structures: (a) GDY_1.5Å_p3%; (b) GDY_1.5Å_p7%; (c) GDY_2.1Å; and (d) GDY_2.5Å. The first two membranes have different porosities of 3% and 7%, respectively. (eh) The criterion for the definition of the pore diameter in each GDY membranes.
Figure 1. Membrane models. Atomic structures of graphdiyne (GDY) membranes with different pore structures: (a) GDY_1.5Å_p3%; (b) GDY_1.5Å_p7%; (c) GDY_2.1Å; and (d) GDY_2.5Å. The first two membranes have different porosities of 3% and 7%, respectively. (eh) The criterion for the definition of the pore diameter in each GDY membranes.
Membranes 10 00286 g001
Figure 2. Simulation system of the equimolar binary mixture of H2/CO2 permeating through the GDY_2.1Å membrane. Atom: C (Cyan); O (red); H (white).
Figure 2. Simulation system of the equimolar binary mixture of H2/CO2 permeating through the GDY_2.1Å membrane. Atom: C (Cyan); O (red); H (white).
Membranes 10 00286 g002
Figure 3. Pure gas permeation. (a) The time evolution of permeated H2 molecules; (b) The permeance of H2 through different GDYs; (c) The permeation of four gases in GDY_2.5Å.
Figure 3. Pure gas permeation. (a) The time evolution of permeated H2 molecules; (b) The permeance of H2 through different GDYs; (c) The permeation of four gases in GDY_2.5Å.
Membranes 10 00286 g003
Figure 4. DFT calculations of gases crossing the membrane of GDY_2.1Å. (a) Minimum energy pathways of four gases. (b) Ideal selectivities of H2 over N2, CO2, and CH4 as functions of temperature.
Figure 4. DFT calculations of gases crossing the membrane of GDY_2.1Å. (a) Minimum energy pathways of four gases. (b) Ideal selectivities of H2 over N2, CO2, and CH4 as functions of temperature.
Membranes 10 00286 g004
Figure 5. H2 permeation of three equimolar binary mixtures crossing the GDY_2.1Å. The final snapshots: (a) H2/N2, (b) H2/CH4, and (c) H2/CO2. Blue: N; cyan: C; red: O; white: H. (d) The H2 permeance of different binary mixtures.
Figure 5. H2 permeation of three equimolar binary mixtures crossing the GDY_2.1Å. The final snapshots: (a) H2/N2, (b) H2/CH4, and (c) H2/CO2. Blue: N; cyan: C; red: O; white: H. (d) The H2 permeance of different binary mixtures.
Membranes 10 00286 g005
Figure 6. Blocking effect on the surface of GDY_2.1Å. (a) Radial distribution function (RDF) of gases around the membrane. Density contours of the impermeable gases on the surface: (b) N2 (H2/N2); (c) CH4 (H2/CH4); (d) CO2 (H2/CO2). The unit of density (Nw/uc) is 1/(1.25Å3).
Figure 6. Blocking effect on the surface of GDY_2.1Å. (a) Radial distribution function (RDF) of gases around the membrane. Density contours of the impermeable gases on the surface: (b) N2 (H2/N2); (c) CH4 (H2/CH4); (d) CO2 (H2/CO2). The unit of density (Nw/uc) is 1/(1.25Å3).
Membranes 10 00286 g006
Figure 7. Permeation behavior of H2 in binary mixtures. (a) The number distributions along the z-direction in the last 10 ns. The membrane is located at the green dotted line; (b) The mean square displacement (MSD) curves of H2 crossing the GDY_2.1Å membrane. (Inset) The region within 0.5 nm of the membrane surface for MSD calculation.
Figure 7. Permeation behavior of H2 in binary mixtures. (a) The number distributions along the z-direction in the last 10 ns. The membrane is located at the green dotted line; (b) The mean square displacement (MSD) curves of H2 crossing the GDY_2.1Å membrane. (Inset) The region within 0.5 nm of the membrane surface for MSD calculation.
Membranes 10 00286 g007
Figure 8. The effect of the surface charges on gas permeation. (a) The distribution of charges on the network of GDY_2.1Å; (b) H2 permeance of the binary mixture of H2/CO2 as a function of surface charges. (Inset) The time evolution of permeated H2 molecules through the charged GDYs.
Figure 8. The effect of the surface charges on gas permeation. (a) The distribution of charges on the network of GDY_2.1Å; (b) H2 permeance of the binary mixture of H2/CO2 as a function of surface charges. (Inset) The time evolution of permeated H2 molecules through the charged GDYs.
Membranes 10 00286 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, Q.; Cheng, L.; Liu, G. Enhanced Selective Hydrogen Permeation through Graphdiyne Membrane: A Theoretical Study. Membranes 2020, 10, 286. https://doi.org/10.3390/membranes10100286

AMA Style

Liu Q, Cheng L, Liu G. Enhanced Selective Hydrogen Permeation through Graphdiyne Membrane: A Theoretical Study. Membranes. 2020; 10(10):286. https://doi.org/10.3390/membranes10100286

Chicago/Turabian Style

Liu, Quan, Long Cheng, and Gongping Liu. 2020. "Enhanced Selective Hydrogen Permeation through Graphdiyne Membrane: A Theoretical Study" Membranes 10, no. 10: 286. https://doi.org/10.3390/membranes10100286

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