Min
Gao‡
a,
Alston J.
Misquitta
a,
Chenxing
Yang§
a,
Ilian T.
Todorov
b,
Andreas
Mutter¶
c and
Martin T.
Dove
*a
aSchool of Physics and Astronomy, and Materials Research Institute, Queen Mary University of London, Mile End Road, London, E1 4NS, UK. E-mail: martin.dove@qmul.ac.uk
bScientific Computing Department, STFC Daresbury Laboratory, Sci-Tech Daresbury, Keckwick Lane, Daresbury, Cheshire WA4 4AD, UK
cDepartment of Earth Sciences, University of Cambridge, Downing Street, Cambridge, CB2 3EQ, UK
First published on 19th July 2017
We report a study of the absorption of CO2 into a number of zinc imidazolate framework structures, and subsequent desorption, using the molecular dynamics simulation method with force fields partly developed by ourselves. The simulations primarily give results concerning the mechanism of CO2 absorption under conditions likely to be found in industrial waste gas streams. In particular we compare the rate of uptake of CO2 for different ZIFs. We also show that it is possible to observe desorption by reduced pressure and high temperature. These characteristics confirm that ZIFs might be suitable for CO2 absorption within an industrial capture and sequestration process.
Design, System, ApplicationZIFs have been proposed as materials for CO2 gas capture. We have developed a new simulation model and strategy to study CO2 absorption in such materials, including a new model for CO2 that works in both gas and condensed phases. This method can be adopted to study all kinds of gas absorption in ZIFs and other metal–organic-frameworks. It can also be used to study gas selectivity of MOF membranes. CO2 absorption and de-absorption in a number of ZIFs have been studied. The results help us to understand the mechanism of the interaction between CO2 and the ZIF structure, and to predict the absorption storage for similar structures. We show that ZIFs have great potential for CO2 capture. For future application, a) our work shows that ZIFs have great potential for CO2 capture; b) our method can be more widely used for many future applications. It can speed up studies of gas absorption, gas selection and storage by screening. This helps choosing appropriate MOFs for particular purposes such as toxic gas absorption in power plant, and hydrogen storage and transport for clean power, etc. |
In this paper we study the absorption and desorption of CO2 in one class of MOFs, namely the materials with the generic name zeolitic imidazolate frameworks (ZIFs).5 These have the chemical formula Zn(im)2, where im = C3N2H3− or a related ligand. These materials form as networks of zinc cations linked by the im ligands. The angle subtended by the metal–im–metal vectors in many ZIFs (ca. 145°) is very similar to the Si–O–Si angle in silica, SiO2, meaning that the networks are able to form structures whose topologies are either direct analogues of, or related to, siliceous zeolites. Furthermore, the structures formed by these materials can be tailored by changing the metal cation, tuning the exact imidazolate linkage ligands,6 and growing from different solvents. This has led to at least 100 types of ZIFs being found.6,7 Their various porous structures offer the prospect of being exploited for applications in areas such as catalysis, storing gases such as H2, and capture of gases such as CO2.6,8–13
Our approach is to use the molecular dynamics simulation method using empirical force fields, focussing on the ZIFs identified in the first report of these materials.6 In the work reported here we started with a pristine slab of a ZIF in contact with CO2 gas, and applied an appropriate temperature and pressure. The pressure had the effect of pushing the CO2 molecules into the ZIF slab. In this paper we report our observations on the processes by which the CO2 can enter the ZIF slabs from the gas phase, and how the molecules move within the ZIF. Here we will also report the structures formed once maximum capacity has been more-or-less reached. We have performed simulations at different pressures, and we compare the results for the different ZIF systems. We have also investigated the desorption process induced by reduced pressure and increased temperature, but because of technical issues this proved to be rather harder.
In the next section we describe the methods used in this work, including details of the force fields we have developed, and the way we have used the molecular dynamics method in this work. Our simulation results are reported in the subsequent sections. The first set of results concerns the absorption of CO2 into slabs of a range of different ZIFs, looking at both the time scales for absorption and then the CO2 density profiles through the slab. Results are presented for two gas pressures. In order to understand the absorption process the section after considers in more detail the interactions between imidazolate ligands and CO2 molecules. In the penultimate section we discuss the process of de-absorption with reduced pressure and increased temperature, with some conclusions presented in the final section. The ESI† presents more results than can be accommodated in the main paper.
Ab initio DFT calculations for small clusters of zinc cations and ligands were performed using the quantum chemistry program NWChem16,17 with the standard PBE functional,18 and using either aug-cc-pVDZ or aug-cc-pVTZ basis sets depending on system size and following convergence tests. We used the distributed multipole analysis (DMA) method19 as performed by the program CamCASP20 to calculate atomic multipole moments from the electronic wave functions, from which we obtained values for atomic point charges using program MULFIT.21,22
The new interatomic potentials for our force fields were obtained by fitting to the energies for a large suite of atomic configurations obtained from ab initio calculations, also performed using the program NWChem,16,17 but this time using the Møller–Plesset perturbation theory MP2 approximation. As for the DFT calculations, we used aug-cc-pVDZ and aug-cc-pVTZ basis sets. Values for the parameters in the force field equations (see immediately below) were obtained using the fitting procedure with the code GULP.23
E(r) = ε(exp(2α(r − r0)) − 2exp(α(r − r0))) | (1) |
E(θ) = (k/2)(θ − θ0)2 | (2) |
E(ϕ) = A(1 − cosϕ) | (3) |
For non-bonded interactions, such as between the atoms of different imidazolate ligands for example, we used the standard Buckingham form to represent both the long-range dispersion interaction and the short-range repulsion:
E(rij) = −(Cij/rij6) + Bijexp(−rij/ρij) | (4) |
For interactions between the imidazolate ligands we used the Buckingham potentials (eqn (4)), with parameters taken from the work of Williams.24
Values for the parameters in the Buckingham potential were obtained by fitting to a mixture of data. We used a suite of ab initio energies calculated for a large number of configurations of pairs of molecules of orientations as shown in Fig. 1. These ab initio energies were calculated using the MP2 method with aug-cc-pVTZ basis sets. We also made use of the crystal structure and phonon dispersion curves of solid CO2 in the fitting procedure, which was carried out using the code GULP.23 We also used Grand Canonical Monte Carlo (GCMC) simulations of the fluid phase diagram to help tune and constrain parameter models.
By this route we were able to tune values for all the parameters. The final set of parameters for our new CO2 model are given in Table 1. Part of this work was briefly described in an earlier paper on fluid CO2,26 but here parameter values have been slightly refined.
Atom pair | B (kJ mol−1) | ρ (Å) | C (kJ mol−1 Å−6) |
---|---|---|---|
C–O | 190909 | 0.2637 | 1216.21 |
O–O | 203567 | 0.2659 | 2149.19 |
C–C | 108347 | 0.2778 | 0.0 |
A model for CO2 with non-rigid molecules are also developed and tested. A Morse potential was used for the stretching of the C–O bond, and a potential of form E(ϕ) = K(1 + cosθ) was used to describe flexing of the O–C–O angle θ; parameter values were tuned to give best agreement with the oft-reported frequencies of the internal molecular vibrations. The highest frequency vibration of the CO2 molecule, namely the asymmetric C–O stretching mode, has a frequency of 70.5 Hz, which is much higher than CO2 intermolecular vibration frequency (4 THz) as shown in Fig. 2, and higher than the highest network flexing frequency of the ZIFs of 14 THz.14 If we scale the timestep by the ratio of frequencies, 14/70.5 ≃ 0.2, we would require computational resources five times greater than we used for the rigid molecule model; given the very large cost of computer time incurred in this work, such an overhead would need significant justification. In the event our tests on the fluid phase diagram showed that the rigid and flexible molecule models gave the same results. Hence only the rigid body model was used here.
Fig. 2 Phonon dispersion curves of CO2. The diagram on the left shows the calculation of our model (continuous curves) and the diagram on the right shows the corresponding calculations using the Potoff and Siepmann model.27 The filled black circles are experiment data from inelastic neutron scattering.29 |
Atom pair | ε (K) | σ (Å) |
---|---|---|
C–C | 27.0 | 2.80 |
O–O | 79.0 | 3.05 |
C–O | 46.18 | 2.92 |
The crystal structure was taken from the X-ray refinement of Simon and Peters.28 CO2 has Pc3 space group symmetry, with the lattice parameter a = 5.624 ± 0.002 Å. The C–O distance is 1.15 Å. Our structure optimisation performed using GULP gave a lattice parameter 5.624 Å in perfect agreement with the experimental result, whereas the optimised lattice parameter for the Potoff and Siepmann model is 4.998 Å with 11% difference from the experiment. Typically we expect a good empirical model to have an accuracy for the lattice parameters of up to 2% (perfect agreement with experimental data should not be taken too seriously), and so we feel that the Potoff and Siepmann model has too poor level of agreement with experimental data in the crystalline phase.
To test the CO2 model of the solid phase further, the phonon dispersion curves have been calculated using GULP23 and compared with the inelastic neutron scattering data of Dolling et al.29 (phonon measurements were performed for wave vectors along the [1,1,1] direction). Calculated phonon dispersion curves for both the model developed here and from that of Potoff and Siepmann are compared with the experimental data in Fig. 2. The calculations were performed with the relaxed structure. The experimental results were obtained by inelastic neutron scattering. From the figure we can see that our calculated dispersion curves show a good agreement with the experimental data. The results for the two acoustic modes and the range of optical phonon frequencies at wave vectors [0,0,0] and of our model are close to the experimental data. On the other hand, the phonon frequencies calculated using the Potoff and Siepmann model are systematically too high.
GCMC simulations were performed for a range of temperatures from 200 K to 300 K, using the constant pressure and temperature Gibbs ensemble. The temperature and pressure coexistence lines are shown in Fig. 3 together with the density at the co-existence line. The simulation results show a satisfying agreement with the experiments. Detailed results for the vapour–liquid coexistence point are given in Table 3 and compared with results for the Potoff and Siepmann model. The accuracy of our model is comparable with that of Potoff and Siepmann for the fluid phase diagram, and has the advantage of also handling the crystalline phase much better.
Fig. 3 CO2 vapour–liquid coexistence line. The black dot and star is the critical point from experiment30 and our simulation respectively. |
Temperature (K) | Pressure (bar) | Vapour density (g ml−1) | |
---|---|---|---|
Experiment | 304.1 | 73.8 | 0.4676 |
Our model | 308.4 | 80.8 | 0.4670 |
PS model | 306.2 | 77.7 | 0.4649 |
Atom pair | B (kJ mol−1) | ρ (Å) | C (kJ mol−1 Å−6) |
---|---|---|---|
CCO2–CZIF | 106996.4 | 0.278 | 1487.38 |
CCO2–H | 60683 | 0.2793 | 713.33 |
CCO2–N | 103366 | 0.282 | 1473.04 |
CCO2–Zn | 1417347 | 0.290 | 38015.38 |
O–CZIF | 163495 | 0.265 | 1394.10 |
O–H | 57298 | 0.266 | 613.98 |
O–N | 157948 | 0.269 | 1380.67 |
O–Zn | 274353 | 0.275 | 11435.99 |
The simulation configurations consisted of slabs of the ZIF structures with three-dimensional periodic boundaries, with the space between the slabs filled with CO2 molecules. The slabs were created always with the crystal [001] axis normal to slab surface, cutting the Zn–N bonds at a position designed to minimise the number of bonds to be cut. We always therefore align the normal to the slab with the sample Z axis, which will be denoted as such in the rest of this paper. The thickness of the slab was chosen to contain 10 cavities, and the lateral dimensions were chosen to be around twice the thickness. The thickness thus depended on the specific ZIF, and varied from 75 Å for ZIF-zni to 211 Å for ZIF-6. To make the system electrically neutral, we attached a hydrogen atom to each open nitrogen atom, as shown in Fig. 5.
The simulations were performed under pressures and temperatures corresponding to the conditions that might be found in a typical industrial process. The plan with the simulations was that the pressure would force the CO2 molecules to enter the ZIF slab. This would result in the volume occupied by the gas to be continuously decreasing, until almost all the molecules in that particular stage of the simulation were inside the slab. Typically the height of the gas column at the start of the simulation was around 800–1000 Å. After one simulation, which will have ended when most of the CO2 molecules had entered the slab, we added a new column of CO2 on top of the ZIF slabs once the CO2 molecules get into the slab, as shown in Fig. 6. The left pane of the figure shows the start of the first stage simulation, and the right pane shows start of the second stage simulation with newly added CO2 position above the slab and with CO2 absorbed within the slab from the first stage simulation. We define one complete absorption process as the suite of simulations to the saturation state where no more molecules are easily absorbed. For one complete absorption process, several stage simulations are needed. The onset to saturation was most easily monitored through the rate of change of volume with time.
To perform the task of adding CO2 molecules to the new DL_POLY configuration file and increase the dimension of the simulation box at each new stage during the absorption process we used a self-written Python script. During the simulation, the coordinates of the system shift along the Z axis when a large amount of CO2 been pushed in, and this shift cannot be predicted, which means that the boundary of the simulation box might cut through the middle of the solid absorbent or the CO2 gas. Due to this shift, any bonds that cross the periodic boundary along the normal axis would be elongated when the length of this axis be increased. Our python script was also used to divide the CO2 gas by a given plane normal to the Z axis and reset this plane as the new periodic boundary.
For the data analysis, we wrote one Python script for extracting the output data from the file generated by DL_POLY that contains a set of data for each time step to be recorded (called STATIS), and another Python script to calculate the atom density profiles along the three axes. The visualisation software CrystalMaker33 was used to produce snapshot images and movies to provide additional information.
In addition to the absorption simulation, reverse desorption simulations were also performed. These started with the configurations for which absorption had reached their equilibrium state, but the configuration was set to atmospheric pressure.
It is useful to note that the large volume changes that occur with out approach to study absorption and desorption are not really compatible with the way that the architecture of the MD software is optimised for parallel computing architectures. In particular it proved impossible to run desorption simulations with DL_POLY4, so for these simulations it was necessary to use the older DL_POLY Classic software, which uses a different parallelisation strategy more suitable for smaller samples. The computational overhead limited the extent to which it was possible to study the desorption process, so the results presented here are rather more limited than for the studies of absorption.
The absorption results are summarised in Table 5, which characterises each ZIF in terms of the number of atoms in each simulation and the density of Zn atoms, and then comparing the number of CO2 molecules absorbed in the simulation and the density of absorbed CO2 molecules. For each case Table 5 also gives the time required to reach saturation in the simulation. The pore (or channel) density is inversely proportional to the density of Zn atoms, with a higher Zn density indicating a smaller pore structure and hence smaller pore surface areas. We might have expected to find a higher CO2 gas capacity in the ZIFs with higher pore surface areas, but the simulation results in Table 5 show no simple link between the Zn density and CO2 uptakes. The results implies that the uptake capacity is decided by more factors rather than just the Zn density.
N ZIF | N CO2 | D Zn | D CO2 | Time (ps) | |
---|---|---|---|---|---|
ZIF-zni | 28008 | 1455 | 7.26 | 6.64 | 4859 |
ZIF-2 | 86112 | 9781 | 5.35 | 10.64 | 6656 |
ZIF-3 | 80896 | 11464 | 4.30 | 10.69 | 1369 |
ZIF-4 | 80896 | 9241 | 5.22 | 10.45 | 4029 |
ZIF-6 | 180992 | 25487 | 4.12 | 9.97 | 8205 |
ZIF-8 | 147584 | 3553 | 4.06 | 2.30 | 21085 |
ZIF-10 | 230496 | 34701 | 4.18 | 10.86 | 16508 |
We remark that the density of the CO2 gas outside the slab in our simulations after saturation with pressure of 25 bar and temperature of 200 °C is approximately 34 kg m−3. This is higher than both the experiment value of 29 kg m−3 at the same condition, which in turn is consistent with our own MD simulations of the gas state showing a density of 28.5 kg m−3 at the same conditions. We speculate that the increased density is associated with attractions due to the surfaces, although if this is the explanation it was not possible to test this because the whole sample size was not large enough to observe the expected gradient in the density in the direction between the surface of one slab and the closest surface of its periodic image.
The patterns of CO2 absorbed in different ZIFs also reflect the structures of pores and channels. This is shown for example of ZIF-4 in Fig. 9. The observation of the pattern of pores reflects the high degree of filling that is being achieved in the simulation. The patterns for absorption in other ZIFs are given in the ESI.†
The change in volume of the system (slab plus gas) in different ZIFs under increased pressure is illustrated in Fig. 10, where we show the effect of switching instantaneously from 25 bar to 30 bar. It can be seen that the volumes decreased rapidly under the higher pressures. This shrinking process will in part be associated with compression of the gas. However, there is also a higher degree of absorption, as shown by the results given in Table 6. They show that the absorption uptake capacity of most of the ZIFs (ZIF-zni, ZIF-2, ZIF-4, ZIF-6 and ZIF-10) under 30 bar is slightly higher than at 25 bar, by just under 5%. The increase in uptake is much larger in ZIF-3 and ZIF-8 slabs, of 8% for ZIF-3 and 16% for ZIF-8. This mechanism of this high increase can be revealed by the CO2 density profiles within the ZIF. The density profile of absorbed CO2 in the ZIF-3 slab under 30 bar is shown in Fig. 11. It shows that the density peaks for 30 bar are higher than the ones under 25 bar. This shows simply that more CO2 molecules have been pushed in, and that they tend to cluster together in the same sites.
Fig. 10 Volume change of CO2 absorption in ZIFs. The black lines represents the adsorption under 25 bar and the red lines represents the adsorption under 30 bar. |
N CO2 (25 bar) | N CO2 (30 bar) | D CO2 (25 bar) | D CO2 (30 bar) | P increase | |
---|---|---|---|---|---|
Units: | mol ml−1/kg m−3 | mol ml−1/kg m−3 | |||
ZIF-zni | 1448 | 1502 | 6.64/278.96 | 6.87/301.97 | 3.43% |
ZIF-2 | 9739 | 10013 | 10.64/468.12 | 10.90/479.55 | 2.42% |
ZIF-3 | 11458 | 12403 | 10.69/470.21 | 11.52/506.82 | 7.84% |
ZIF-4 | 9222 | 9462 | 10.45/459.75 | 10.68/469.87 | 2.22% |
ZIF-6 | 25371 | 26561 | 9.97/438.63 | 10.41/547.99 | 4.47% |
ZIF-8 | 3552 | 4066 | 2.30/101.19 | 2.63/115.7071 | 14.53% |
ZIF-10 | 34630 | 35187 | 10.86/477.79 | 11.03/485.27 | 1.62% |
Fig. 11 Absorbed CO2 density profile in ZIF-3 slab. The black line represents the adsorption profile for 25 bar, the red lines represents the adsorption profile for 30 bar. |
We have also made a animation of CO2 in ZIF-zni at the saturation state, using CrystalMaker, which is provided within the ESI.† The animation shows that the CO2 molecules in saturated ZIF-zni move back and forth with the translational and rotational vibrations of the ligands.
The way to handle the desorption simulation is basically a reverse process of absorption simulation. Generally, the simulations have been performed in separate stages similar to absorption simulations. When the gas was released to a certain amount, typically an expansion in the direction of the normal to the slab by a factor of about 4 or 5, the gas molecules were removed from the simulation manually and the volume adjusted accordingly. This new configuration was then the starting point for the subsequent simulation.
The desorption process was initially simulated using the same MD code DL_POLY used for the absorption simulations, but we found that this code gave some problems associated with its use of high-performance computing architectures. We found that the same problems did not arise using a previous version, DL_POLY classic (version Classic 1.9), but this has the disadvantage of not being optimised for large systems in the same way that Version 4 has been. Thus we used version 4.06.1.1 for the initial desorption simulations, and then we switched to DL_POLY classic for further desorption runs. This strategy, however, only allowed us to perform short period simulations of the desorption process.
Fig. 13 CO2 desorption dynamics in ZIFs at 200 °C under 1 bar. The configuration of the first run (I run) is taken from the last output files of absorption under 25 bar. |
The decreased percentage of absorbed CO2 molecules are shown in Table 7. The Table shows that big portions of the gas have been released in ZIF-2, ZIF-3 and ZIF-4 already. It takes longer time to release the gas in ZIF-2. However, their trends in Fig. 13 indicates that the systems are still far away from the equilibrium state which means a larger amount of CO2 can still be released from these ZIFs, including ZIF-10.
N CO2 (25 bar) | N CO2 (1 bar) | Time (ps) | Decreased percentage | ||
---|---|---|---|---|---|
ZIF-2 | 9781 | 8710 | 402 | 16.47% | |
ZIF-3 | 11464 | 8547 | 555 | 25.44% | |
ZIF-4 | 9241 | 6588 | 903 | 28.71% | |
ZIF-10 | 34701 | 32978 | 462 | 5% |
We have also performed a simulation of the desorption process at the higher temperature of 400 °C, with results shown in the last image in Fig. 13. It took approximately 66 ps at 200 °C to expand to the same volume, but only 42 ps at 400 °C. The number of CO2 molecules that came out of the slab further showed that the gas molecules are released quicker at higher temperature. While only 684 CO2 molecules have been released at 200 °C after 66 ps, in the 42 ps to give the same volume the slab released three times the number of molecules, namely 2294 molecules. This means that the absorbed CO2 molecules can be released significantly faster by heating up the system.
Based on these model force fields, a set of systematic MD simulations on CO2 absorption process in ZIF slabs have been presented. The simulations on CO2 absorption in ZIFs have been performed with environments close to those found in industrial power plants. The simulations have produced a large quantity of data, some of which have been reported in this paper but with many results presented in the ESI.† The simulations have given insights into the process of absorption, showing both similarities and differences between the behaviour of different ZIF systems. It is found that it is usually possible to achieve a high filling of the porous structures, as monitored both by calculations of density profiles and visualisations. It has been found that the preferable absorption sites for CO2 molecules are close to the imidazolate ligands due to the CCO2–NZIF Coulomb attraction.
In addition to the absorption simulation, the desorption processes have also been simulated. The desorption simulation results have shown that the CO2 molecules can be taken out by decreasing the pressure or increasing the temperature.
All these results point to the real possibility of using porous structures like ZIFs for capture of CO2 in an industrial environment as part of a strategy to reduce the release of CO2 into the Earth's atmosphere. In a following paper we will discuss some of the details of the processes at the atomic scale, but here our headline results concern the potential of such materials for this application, but with the caveat that different members of the general family of materials may show very different absorption capacity.
In future work it would be useful to work with gas streams containing other molecules, such as N2 and H2O. This will require an extension of our force field model, ideally using the approaches discussed in this paper.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c7me00034k |
‡ Present address: China Spallation Neutron Source, Institute of High Energy Physics, Chinese Academy of Sciences, 1 Zhongziyuan Road, Dongguan 523803, People's Republic of China, and, Dongguan Institute of Neutron Science, 1 Zhongziyuan Road, Dongguan, 523808, People's Republic of China; E-mail: E-mail: gaomin@ihep.ac.cn |
§ Present address: Institute of Natural Sciences, and, Department of Physics and Astronomy, Shanghai Jiao Tong University, Shanghai 200240, People's Republic of China. |
¶ Present address: Fachbereich Physik, Erwin-Schrödinger-Strasse, 67663 Kaiserslautern, Germany. |
This journal is © The Royal Society of Chemistry 2017 |