Introduction

A current major challenge in science and technology is the development of low cost materials with large CO2 capture and storage (CCS)1 capacity, retention ability and sorption selectivity2. Important factors for the choice of material in this context include its availability, environmental friendliness, non-toxicity, a low level of greenhouse gas emission during processing, material stability, production cost, CO2 storage capacity, energy of adsorption/desorption, sorbent regeneration, sorption kinetics and capacity per volume or per mass of host material. Many materials are being investigated including zeolites3, metal organic frameworks (MOFs)4, functionalized polymers5,6, activated carbons7 and nano-silicate clay minerals8,9,10. Also hybrid solutions using the best characteristics among all these absorbents are much investigated11. This work presents a quantitative study of the capture of CO2 by Fluorohectorite clay - a stable and low cost material - with different interlayer cations. Fluorohectorite clay is a modified hectorite clay, which is a species belonging to the smectite group12, classified as a 2:1 layered silicates, with layer periodicity of approximately 1 nm. The periodic structure includes an interlayer space where a cation is lodged, bonding the structure and balancing the overall charges. The group of smectites, commonly addressed in short as clays, also comprise more popular species like montmorillonite clays. The synthetic fluorohectorite, studied in the present work, has been demonstrated, in several publications by our group, to be a representative and clean model system of nano-silicate smectite clay minerals13. Synthetic nano-silicate clays contain significantly fewer impurities (e.g. carbonates, (hydr)oxides, silica, and organic matter) than natural clays and show a more homogenous charge distribution than their natural counterparts14, leading to well-defined intercalation states15,16. Intercalation of water in nano-silicate smectite clays occurs naturally and has been extensively studied with a wide range of techniques, among them neutron17,18 and X-ray scattering13,19. Recent experiments and simulations have shown that also CO2 can intercalate in smectite clays, both in supercritical20, and in gaseous/liquid form21. We have previously demonstrated22 that under certain conditions of pressure and temperature, fluorohectorite clays can capture a large amount of CO2 depending on the type of interlayer cation, and that Nickel-fluorohectorite clay (NiFh), in particular, will retain CO2 up to a temperature of 35 °C at ambient pressure23. The captured CO2 can subsequently be released by heating above this temperature. These conditions are highly relevant for mapping out and understanding the mechanisms involved in CO2 capture and retention by nano-silicates, either in geological formations, or in technological CO2 absorbent materials.

Results and Discussion

We have utilized an accurately calibrated custom-made Sieverts apparatus for the quantification of CO2 storage capacity of the nano-silicate clay minerals. The setup is described in detail in the Methods section. The high-pressure CO2 uptake at room temperature (Fig. 1a) shows a maximum CO2 intercalation of 28% in weight for NiFh clay at the final pressure of 55 bar. This value is equivalent to 6.45 mmol of CO2 per gram of NiFh, which corresponds to 0.795 ton of CO2 per m3 of clay at 55 bar and room temperature, given the crystallographic density of the clay (2.8 g/ml). This result demonstrates that NiFh is able to intercalate one of the largest amounts of CO2 per volume of material of all porous materials reported in literature. Commercially available zeolite3 13X can adsorb 0.35 ton/m3, the mesoporous carbon7 MPPY-4800 has been reported to uptake 0.93 ton/m3, whereas MOF-2104 exhibits a volumetric capacity of 0.72 ton/m3 of CO2 per volume of host material. Nevertheless, MOFs report the highest gravimetric capacity records because of their low density. Table 1 includes a comparison of exceptional results on CO2 uptake - showing both volumetric and gravimetric values - and specifically our results on high volumetric capacity clays, calling the attention to the diversity of characteristics among all materials that could be combined for a viable carbon capture and storage solution.

Figure 1
figure 1

(a) Total uptake of CO2 into Fluorohectorite nano-silicate clay minerals determined for each pressure step at room temperature; (b) Incremental uptake of CO2 into Fluorohectorite clays for each pressure step.

Table 1 CO2 intercalation values and clay parameters.

The other clay types investigated here, NaFh and LiFh, are also found to be high capacity storage materials, with intercalation capacity of 21 wt % (corresponding to 0.58 ton/m3 or 4.7 mmol/g) and 16 wt % (corresponding to 0.46 ton/m3 or 3.7 mmol/g), both at a final pressure of 53 bar respectively. These numbers are higher than those found in a theoretical study on other clays9, for example, Na-Montmorilonite with a capacity of 1.5 mmol/g for a basal interlayer distance of 12 Å, which is comparable with the interlayer distance of our loaded samples as shown in Table 2. The volumetric capacity of LiFh was already studied by us and published elsewhere23. Here we report new results on LiFh for a comparison using the same experimental protocol (dehydration, incubation time and incremental steps in pressure) as used for NiFh. The results for NiFh are original and demonstrate a significant development forward regarding CO2 capture by a nano-silicate material.

Table 2 Fluorohectorite clay parameters.

NaFh and LiFh present very similar behavior in terms of the intercalation of CO2, as shown in Fig. 1. This suggests a similar interaction between the CO2 molecule and the Na+ and Li+ cations. They are both alkali metals of group 1 A with comparable ionic distribution in the nano-silicate clay interlayers. Previous results showed that after the intercalation into Fluorohectorites, CO2 is stable inside the nano-silicate clay material and can be released upon heating23, which indicates the formation of stable cation-CO2 complexes or cation solvation into a CO2 gas phase24. The size and structure of the complex might favor the stability and can be probed by X-ray diffraction (XRD). The XRD results from our previous work23, transcribed in Table 2, show that the interlayer separation occupied by CO2 for NaFh and LiFh increases with the ionic radius of the cation. Na+ has a larger ionic radius than Li+, and the loaded NaFh is found to have a larger interlayer separation than LiFh, which means additional room to accommodate a larger complex or higher coordination number. This could explain our finding that NaFh intercalates more CO2 than LiFh at the investigated pressures.

The total CO2 intercalated (Σ Ui) as a function of pressure is shown in Fig. 1a, while Fig. 1b shows the incremental amount (Ui) of CO2 intercalated at each step in pressure. The latter indicates that the process is driven towards a saturation limit, after which no additional intercalation will be possible. Since the incremental values have not reached zero at the maximum pressure used in this study, it is possible that CO2 intercalation is not yet at the full potential for this material. At this stage, despite presenting already a high uptake of CO2, further investigations into optimization of the CCS capacity of these materials may thus benefit from an extended pressure/temperature protocol as well as exploring the use of functionalized clays such as those prepared according to Breu’s group25. Reuse of the CO2-loaded clay materials is possible after releasing the CO2 by a combination of moderate heating and evacuation. This has been tested for NiFh, NaFh and LiFh (data not shown), with no sign of degradation of the material.

Table 1 shows the results for the CO2 gravimetric and volumetric capacity for the clays studied in the present work, compared with Montmorilonite26, Zeolite 13X3, functionalized polymers5, activated carbons7, MOF–aminoclay composites11 and MOF-2104 data from the literature. The final pressure varies according to the characteristics of the material. Some materials, like the fluorohectorite clays, have low uptake for flue gas but are not saturated at low pressure, allowing storage at high pressure. Table 2 shows the molar weight, number of CO2 molecules per unit cell (N) and the basal interlayer distance (d001) for each clay type.

The molecular formula per unit cell for NiFh, NaFh and LiFh is expressed by

$${{\rm{M}}}_{{\rm{x}}}({{\rm{Mg}}}_{6-{\rm{x}}}{{\rm{Li}}}_{{\rm{x}}}){{\rm{Si}}}_{8}{{\rm{O}}}_{20}{{\rm{F}}}_{4}$$

where x = 1.2 for LiFh (M = Li) and NaFh (M = Na) and x = 0.6 for NiFh (M = Ni)27. The number N is found to be 5 in the case of NiFh. This means more than 8 molecules of CO2 per Ni2+ cation, since NiFh has approximately 0.6 Ni2+ cations per unit cell28. The two other clays, LiFh and NaFh, contain 1.2 cation per unit cell, which gives a value for the ratio N/cation between 2 and 3.

The Fig. 2 shows possible structures of CO2-cation complexes assembled into NaFh and NiFh clay materials, respectively. These structures use as a starting point the crystal structure of water-containing clay, with cell parameters a = 5.2432 Å and b = 9.0870 Å, as published by Kalo et al.8, but here with the water molecules removed and substituted by another set of three atoms in the form of CO2. Our sketch of the CO2 intercalated fluorohectorite is intended to give a plausible picture of how the CO2 molecules could be arranged between the layers upon removal of water molecules. These hypothetical structures were derived by iterative process taking into account the nature of CO2-CO2, and CO2-host intercalations in a most qualitative way, while keeping interatomic distances from becoming too short. This was done without expanding the layer spacing from what it was for water. The three shortest CO2 oxygen-to-host-oxygen distances were 2.31, 2.53 and 2.607 Å, but these can, of course, be readily increased by increasing the layer spacing. We have started with an interlayer distance of 10 Å, yielding a layer periodicity of 15 Å (as in Kalo et al.8 for 2 water layers), which is close to the XRD results in the work of Michels et al.23 and the simulations in the work of Kadoura et al. (2012)9. The configuration of the intercalated complex is expected to be rather dynamic and the number of CO2 molecules around each cation is here estimated to be 2 for NaFh and 8 for NiFh based on the amount of material intercalated. For free Ni2+, the coordination number29 with CO2 can be as high as 6, and in Fig. 2(a) we assume this to be the case for intercalated Ni2+ as well. The binding of the two additional satellite molecules of CO2 could come from interaction with the clay layers.

Figure 2
figure 2

Suggested configuration for CO2-cation complexation inside Fluorohectorite clay interlayer: (a) 8 molecules of CO2 for each Ni2+ cation and (b) 2 molecules of CO2 for each Na+ cation. (The structures were built using the software CrystalMaker®).

In conclusion, we have in this work compared alternative materials with exceptional gravimetric or volumetric capacity for CO2 capture and storage. We have demonstrated that NiFh is able to capture 0.79 ton of CO2 per m3 of clay at 55 bar and ambient temperature. This is one of the highest volumetric capacities of all CO2 sorbent material reported to date. This result, combined with the diversity of characteristics of other high performance materials, could lead to new technological solutions for Carbon Capture and Storage (CCS). The storage capacity for fluorohectorite clays may well be further enhanced beyond the present results by utilizing functionalized clay minerals8. Moreover, it has also been suggested that clay materials can be used in gas separation, in particular, those of greenhouse gases such as methane9 from CO2.

Clay materials such as the synthetic Fluorohectorites investigated here are harmless30 and in addition, their production is expected to be substantially less expensive than other high performance porous synthetic materials considered for CO2 capture, such as MOFs whose upscaling is particularly challenging. Furthermore, clays are also proven to be highly stable when exposed to external mechanical stress or humidity on geological time scales.

Our results show that the type of clay-interlayer cation is critical for CO2 capture, whereas the clay nanolayers provide the large effective surface area responsible for the exceptional storage capacity. The remarkable capture capacity of NiFh is also related to (a) a high charge density, about twice that of Montmorillonite28, and (b) the presence of divalent Ni2+ that frees up additional space for CO2 intercalation, which results in a higher uptake per cation compared with the usual monovalent ions such as Na+. The internal surface area for fluorohectorite and montmorillonite was found to be 3 m2/g and 80 m2/g respectively in the work of Kaviratna et al.28 (using the fluorohectorite sample bought from the same batch as in the present work), which could be an indication that the CO2 uptake, in these two clays, is more dependent on the interlayer separation and interlayer cation than on the internal surface area. We believe that the main interaction mechanism responsible for the high CO2 affinity is direct dipolar electronic interaction with the interlayer cation. The physisorption is probably also, to some degree, influenced by the charge distribution of the clay layered structure.

Other aspects of sorption kinetics are under evaluation and DFT calculations of the intercalate CO2-cation structures are in progress using the initial model presented in Fig. 2 as starting point. This study will be combined with in situ spectroscopic characterization. Experiments using higher pressure, including liquid and supercritical CO2, are also planned. Our Sieverts apparatus allows for temperature measurements which will be important for future determination of heats of adsorption and comparisons with DFT calculations. The present work has been conducted on dehydrated clays; however it is of considerable interest to evaluate the sorption performance in the presence of water and selectivity for CO2 in the presence of CH4 and N2. The current results already make these materials highly interesting candidates for CO2 capture from dried industrial combustion gases.

Methods

We have utilized a custom-made Sieverts apparatus for the quantification of CO2 storage capacity of the clay materials. The instrument has been designed and constructed for minimizing the experimental error and has been accurately calibrated according to procedures described in the literature31. CO2 gas was introduced to the reservoir and equilibrated for one hour, prior to allowing the expansion into the sample chamber and intercalation into the clay powder material. The number of CO2 gas molecules initially in the reservoir, ni = PiVi/(ZiRTi), gas molecules was calculated using the compressibility of a real gas, Z(p,T), as a function of pressure and temperature with parameters from NIST tables as reported elsewhere31. For each pressure increment, the gas uptake was calculated from the difference nf − ni, where nf is the final number of gas molecules expanded to the total volume, Vt (sample chamber + reservoir): nf = Pf(Vt − Vs)/(ZfRTf), where Vs, sample volume, is the mass of sample per its bulk density. Prior to the gas sorption the clay powder was degassed for 20 hours at 120 °C under dynamic vacuum (down to 10−7 mbar), and the subsequent experiments were performed at 25 °C. The synthetic clays Nickel Fluorohectorite (NiFh), Sodium Fluorohectorite (NaFh) and Lithium Fluorohectorite (LiFh) and CO2 gas with purity >99.999% was used. LiFh was purchased from Corning Inc. NiFh and NaFh powder samples were obtained from the LiFh clay following a standard dialysis cation exchange protocol32 with no further analysis. Water intercalation in NaFh and NiFh samples, originated from the same LiFh batch that we have studied here, have been investigated by NMR spectroscopy (NaFh29), and TGA (NiFh19). The cation coordination numbers proposed from these results are in full accordance with the clay layer charge reported by Kaviratna et al.28.

The Sieverts apparatus had a reservoir volume of 16 ml and sample chamber of 13 ml. We have determined the number of CO2 molecules at the reservoir before and after expansion of the reservoir into the sample chamber. The whole process consisted of three stages: (A) degassing of sample and chambers, (B) accumulation of CO2 into the reservoir at a certain incremental step in pressure and, finally, (C) expansion of CO2 into the sample chamber. The total CO2 intercalation was determined as a sum of the intercalation amounts due to each incremental step in pressure. The CO2 intercalation was replicated for testing reproducibility. The amount of powder sample inside the chamber was around 2 g and the bulk density around 0.7 g/ml. The total intercalation period, steps 1 to 3, was uniform around 4 days for all samples. Previous experiments22 showed that the kinetics of intercalation is slower for NaFh compared to NiFh and LiFh, so we chose to expose all samples to the longest period. We expect that the intercalation time might reduce under different conditions, for example with supercritical CO2.