Abstract
Alkaline earth metal oxides are economically attractive, earth-abundant sorbents for CO2 capture. In particular, MgO is a promising CO2 sorbent for applications in the intermediate temperature range (200–400 °C). However, MgO requires the addition of promoters such as alkali metal nitrates or carbonates to overcome the slow kinetics of CO2 sorption. It has been observed experimentally that the addition of K2CO3 increases the CO2 uptake of MgO, yet its role remains poorly understood; this is a critical knowledge gap for the design of more efficient sorbents. In this work, using a combination of in situ XRD and Raman spectroscopy, we gain insight into the CO2 uptake mechanism of K2CO3-promoted MgO, which proceeds in two steps. An initial rapid CO2 uptake (within 1 min) occurs via the formation of a potassium-rich amorphous carbonate phase (K2CO3·xMgCO3·yH2O). This is followed by a slower CO2 capture step, in which the intermediate K-rich phase transforms into a magnesium-rich amorphous carbonate, along with the formation of small amounts of crystalline (hydrated and anhydrous) carbonates such as K2CO3, baylissite (K2Mg(CO3)2·4H2O), and nesquehonite (MgCO3·3H2O). In situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS) analysis reveals that surface hydroxides, which remain on the MgO surface even after a thermal treatment at 720 °C in N2, facilitate the formation of hydrated carbonates at 315 °C in dry CO2. However, the presence of steam during carbonation lowers the CO2 uptake and inhibits the second CO2 uptake step, most likely by stabilizing the K-rich amorphous intermediate. Cyclic CO2 uptake and regeneration experiments reveal the crucial role of a high K2CO3 dispersion within the sorbent for maintaining good cyclic stability.
Keywords: CO2 capture, MgO, K2CO3 , promoter, amorphous carbonate


Introduction
The challenge of reducing atmospheric CO2 emissions to limit the extent of global warming has driven significant interest in developing efficient and cost-effective materials for CO2 capture from large anthropogenic point sources. − In this regard, alkaline earth metal oxides such as MgO are being studied as promising alternatives to the benchmark aqueous amine technology. − This class of sorbents has the potential to be less energy-intensive and avoid the release of toxic gases arising from the oxidative degradation of amines. − MgO-based CO2 sorbents have emerged as promising candidates for industrial-scale CO2 capture due to their high theoretical CO2 uptake capacity of 1.09 gCO2 gMgO –1, low cost, and ability to capture and release CO2 in the intermediate temperature range (200–450 °C). This temperature range is well-suited for precombustion CO2 capture processes. ,, The associated cyclic CO2 capture process of MgO-based CO2 sorbents is based on the reversible reaction of MgO with CO2:
| 1 |
However, bare MgO exhibits slow CO2 uptake kinetics, resulting in CO2 uptake capacities in the range of 0.02–0.04 gCO2 gMgO –1, well below the theoretical maximum. , This limitation has driven research into various strategies to enhance the CO2 uptake performance of MgO, including the use of promoters. ,,
Alkali metal salts, including nitrates (e.g., LiNO3, NaNO3 and KNO3) and carbonates (e.g., Na2CO3 and K2CO3), are among the most effective promoters for MgO-based CO2 sorbents. ,,− For example, MgO promoted with a mixture of (Li, K)NO3 and (Na, K)2CO3, achieves a CO2 uptake of 0.84 g gsorbent –1 (corresponding to a MgO conversion of 95%) after 4 h of carbonation at 325 °C. Despite their efficacy, the mechanisms by which alkali metal salts enhance the CO2 uptake of MgO-based sorbents remain poorly understood and most of the recent mechanistic studies have focused on NaNO3, which is molten under typical carbonation conditions. In contrast, relatively few studies have investigated the mechanisms by which alkali metal carbonates promote the CO2 uptake of MgO, and most of these works have focused on mixtures of alkali metal nitrates and carbonates, such as NaNO3–Na2CO3. , In the presence of a molten alkali metal nitrate, carbonate promoters are generally understood to act as nucleation seeds for MgCO3 formation. ,,− Moreover, many carbonates such as Na2CO3 and CaCO3 readily react with MgO and CO2 to form double carbonates. Compared to alkali metal nitrates, alkali metal carbonates are thermally more stable, less corrosive, and remain in the solid state under typical operating conditions, which makes them more suitable for usage in large-scale reactors such as packed or fluidized beds. ,
Among the alkali metal carbonates studied in the absence of nitrates, the promotion of MgO with K2CO3 has yielded the highest CO2 uptakes. However, there is a very limited body of work on K2CO3-promoted MgO for CO2 capture, and such studies have been conducted under diverse operating conditions, including differences in temperature, gas atmosphere and the amount of K2CO3 used for promotion (Table ). − Overall, the CO2 uptake capacities of K2CO3-promoted MgO are in the range of 0.09–0.18 g gsorbent –1,, well above the typical values of unpromoted MgO (0.02–0.04 g gMgO –1). The increase in the CO2 uptake of MgO upon its promotion with K2CO3 is typically explained by the reaction of K2CO3 and MgO with CO2, leading to the formation of a double carbonate: ,−
| 2 |
1. Overview of K2CO3-Promoted MgO-Based CO2 Sorbents Reported in the Literature.
| Material | CO2 uptake (gCO2/gsorbent) | Operating conditions (all experiments performed at 1 bar) | Reference |
|---|---|---|---|
| MgO-30mol% K2CO3 | 0.033 | 10 vol% H2O; 10 vol% CO2 at 200 °C (carbonation time not specified) | |
| MgO-95mol% K2CO3 | 0.085 | 1 bar CO2; 350 °C for 2 h | |
| MgO-15mol% K2CO3 | 0.119 | 1% CO2, 9% H2O (1 bar) and 60 °C for 2 h | , |
| MgO-15mol% K2CO3 | 0.179 | 1% CO2, 11% H2O and 50 °C for 3 h | |
| MgO-10mol% K2CO3 | 0.144 | 1 bar CO2 at 330 °C for 6 h |
However, the CO2 uptakes observed for K2CO3-promoted MgO (Table ) exceed the expected uptake according to eq both at low and at higher temperatures, indicating that the CO2 uptake by MgO itself was promoted in the presence of K2CO3 (or the double carbonate formed). − Yet, as no detailed characterization of the final product has been reported thus far it is unclear how the (additional) CO2 is captured by MgO, e.g. due to the formation of surface carbonates as in the absence of K2CO3, or due to bulk MgCO3 formation, promoted by K2CO3 (or K2Mg(CO3)2). Furthermore, the reported CO2 uptake of K2CO3-promoted MgO depends strongly on the amount of K2CO3 added. Indeed, the CO2 uptake of a nearly stoichiometric MgO-K2CO3 sorbent (0.085 g gsorbent –1 in CO2 at 350 °C) is lower than that of a sorbent containing only 10 mol % K2CO3 (0.144 g gsorbent –1 in CO2 at 330 °C). The lower CO2 uptake in the presence of a higher loading of K2CO3 could be due to a smaller interfacial area between MgO and K2CO3, or a reduction in the surface area of the sorbent when a large amount of K2CO3 is added. , It is also worth noting that the presence of water vapor strongly affects the CO2 sorption capacity of K2CO3-promoted MgO. At low temperatures (<140 °C), water vapor was found to be essential for CO2 uptake to occur, whereas at higher temperatures (200–320 °C), steam partially inhibits CO2 uptake.
Given the lack of systematic studies on the effect of K2CO3 on the CO2 uptake performance of MgO, this work aims to elucidate how the addition of K2CO3 promotes the CO2 uptake of MgO in the intermediate temperature range. To this end, we investigated the evolution of both amorphous and crystalline carbonates during the carbonation of K2CO3-promoted MgO at 315 °C in CO2 using a combination of in situ X-ray diffraction (XRD), in situ Raman spectroscopy and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS). We observed an initial, rapid formation (<1 min) of a K-rich amorphous carbonate (K2CO3·xMgCO3·yH2O) through the reaction of K2CO3 with MgO and CO2. After approximately 20 min of carbonation, the K-rich amorphous carbonate further reacted with MgO and CO2, resulting in the formation of a Mg-rich amorphous carbonate, explaining the capture of additional CO2. Simultaneously, small amounts of baylissite (K2Mg(CO3)2·4H2O), nesquehonite (MgCO3·3H2O) and K2CO3 slowly crystallized. Furthermore, in situ DRIFTS revealed that the water involved in the formation of the hydrated carbonates originates from H-bonded OH-groups, which remain on the surface of the sorbent even after thermal treatment at 720 °C and recombine to water during carbonation. Interestingly, the addition of steam during carbonation inhibits the second CO2 capture stage, likely by stabilizing the amorphous K-rich intermediate. Finally, we assessed the potential of the sorbent for application in a cyclic CO2 uptake process.
Results and Discussion
CO2 Uptake Performance of K2CO3-Promoted MgO Sorbents
A series of K2CO3-promoted MgO sorbents was prepared by ball-milling MgO with varying amounts of K2CO3 (0, 2.5, 5, and 7.5 mol %). The sorbents are referred to as MgO-xK, where x denotes the nominal K2CO3 loading. Inductively coupled plasma optical emission spectroscopy (ICP-OES) and SEM-EDX analyses (Table S1) confirmed that the molar ratios of Mg:K in the sorbents were consistent with the targeted values. XRD analyses of the as-prepared sorbents (Figure S1) show only Bragg peaks due to MgO for sorbents containing 0, 2.5, and 5 mol % K2CO3, while also peaks due to potassium carbonate sesquihydrate (K2CO3·1.5H2O) were observed in MgO-7.5K. The absence of any XRD peaks corresponding to potassium-containing phases and the presence of a diffuse scattering halo in MgO-2.5K and MgO-5K indicates that potassium was present in an amorphous form. To identify the nature of the amorphous K-containing phase(s), the as-prepared sorbents were interrogated by Raman spectroscopy (Figure S2). In all of the as-prepared K2CO3-promoted sorbents (MgO-2.5K, MgO-5K, and MgO-7.5K), a broad peak at 1082 cm–1 was observed, a peak that is assigned to baylissite, i.e. a hydrated double carbonate of magnesium and potassium (K2Mg(CO3)2·4H2O, see Figure S3). Baylissite readily formed during exposure of the samples to air via the reaction of K2CO3-promoted MgO with water and CO2 from the atmosphere. , Furthermore, a relatively sharp peak at 1061 cm–1, corresponding to the characteristic ν1 peak of K2CO3, was present in MgO-5K and MgO-7.5K. TEM-EDX (Figure a-c) and SEM-EDX (Figure S5) mapping of the as-prepared sorbents revealed that potassium was well-distributed in MgO-2.5K and MgO-5K, although potassium was found to be partially agglomerated in MgO-7.5K. N2 physisorption experiments (Figure S6) indicate that the surface area of the different sorbents decreased with increasing K2CO3 content, resulting in surface areas of 129, 52, 25, and 21 m2 g–1 for MgO-0K, MgO-2.5K, MgO-5 and MgO-7.5K, respectively. The reduction in surface area in the presence of K2CO3 can be attributed to K2CO3 covering the MgO surface, thereby blocking a fraction of the pores of MgO.
1.
STEM-EDX maps of as-synthesized (a) MgO-2.5K, (b) MgO-5K, and (c) MgO-7.5K. (d) CO2 uptake of MgO, MgO-2.5K, MgO-5K, and MgO-7.5K during carbonation at 315 °C in CO2 in a TGA with (e) a magnification of the red dashed area in (d). (f) Stacked XRD patterns of MgO, MgO-2.5K, MgO-5K, and MgO-7.5K obtained after 10 h of carbonation at 315 °C in CO2 in overlay with the reference patterns of nesquehonite (ICSD-91710), baylissite (ICSD-200006), β-K2CO3 (ICSD-662), γ-K2CO3 (ICSD-10191), K2Mg(CO3)2 (ICSD-31295), and MgO (ICSD-52026). (g) Enlarged view of the red dashed area in (f). (h) Visualization of the unit cells of γ-K2CO3, β-K2CO3, MgCO3·3H2O (nesquehonite), K2Mg(CO3)2·4H2O (baylissite), and K2Mg(CO3)2.
The CO2 uptake of the K2CO3-promoted MgO-based sorbents at 315 °C in CO2 was determined using a thermogravimetric analyzer (TGA) (Figure d,e). Prior to carbonation, the sorbents were pretreated at 450 °C in N2 for 30 min to remove adsorbed CO2 and H2O. In the absence of K2CO3, the CO2 uptake of MgO was very low, viz. 0.013 gCO2 gsorbent –1 (weight increase by 1.3 wt.%) after 10 h of carbonation (Figure d). The presence of K2CO3 enhanced substantially the CO2 uptake, yielding CO2 uptakes of 4.4 wt % (0.044 gCO2 gsorbent –1) for MgO-2.5K, 11.4 wt % (0.114 gCO2 gsorbent –1) for MgO-5K and 10.2 wt % (0.102 gCO2 gsorbent –1) for MgO-7.5K. All K2CO3-promoted sorbents showed a fast CO2 uptake at the onset of the carbonation reaction with similar CO2 uptake kinetics (Figure e). Previous studies have attributed this very rapid, initial CO2 uptake to the formation of K2Mg(CO3)2 according to eq . ,,, The amount of CO2 captured in this first, rapid CO2 uptake step (reaction stage I) correlated indeed with the K2CO3 content in the sorbent (when also considering that a small amount of CO2 is captured in the form of surface carbonates on MgO), see Table S2.
Assuming that CO2 is captured through the formation of the double carbonate (K2Mg(CO3)2) during reaction stage I, the results in Table S2 show that MgO-5K reached full conversion of K2CO3. In contrast, a fraction of K2CO3 remained unreacted in MgO-7.5K (full conversion would yield an uptake of 0.069 gCO2 gsorbent –1, but only 0.057 gCO2 gsorbent –1 was observed). This difference is most likely due to a poorer dispersion of K2CO3 in MgO-7.5K. Interestingly, in MgO-2.5K the CO2 uptake (0.035 gCO2 gsorbent –1) exceeded the amount expected for the full conversion of K2CO3 into K2Mg(CO3)2 (0.026 gCO2 gsorbent –1), indicating that the uptake cannot be explained by K2Mg(CO3)2 formation alone, but instead points to the formation of an addtional phase. Interestingly, for MgO-5K and MgO-7.5K, a second, slower CO2 capture stage was observed. A similar two-step CO2 uptake behavior was also reported by Kwak et al. for MgO promoted with 10 mol % K2CO3 at 330 °C in CO2. However, the authors could not identify the formed phases based on their ex situ XRD or solid state NMR analyses.
To identify the phases formed during the carbonation of K2CO3-promoted MgO, XRD analysis of the carbonated sorbents was performed (Figure 1 f,g). For MgO-0K, no crystalline carbonate phases were detected, indicating that the small amount of CO2 captured was due to the formation of surface carbonates on MgO, consistent with previous studies and confirmed by DRIFTS experiments (Figure S16). ,, The XRD pattern of MgO-2.5K shows no sharp reflections; instead, the presence of diffuse X-ray scattering was observed as highlighted in Figure g. The XRD patterns for MgO-5K and MgO-7.5K revealed several small peaks in addition to a broad amorphous background, indicating the presence of several crystalline carbonate phases, including MgCO3·3H2O (nesquehonite), K2Mg(CO3)2·4H2O (baylissite), K2Mg(CO3)2 and K2CO3 (β- and γ-type), in addition to one or multiple amorphous phases. The unit cells of the crystalline carbonates present in the carbonated sorbents are visualized in Figure 1 h and more details on the crystal structures are provided in Table S3. Interestingly, only a small amount of K2Mg(CO3)2 was detected by XRD in MgO-5K and MgO-7.5K. Additional K2Mg(CO3)2 may be present in an amorphous form, may have transformed into baylissite by reacting with water upon exposure to air during the XRD measurements, or a different K2CO3-containing phase may have formed.
To obtain more structural insight into the nature of the amorphous phase(s) formed during carbonation, MgO-2.5K and MgO-5K were probed by Raman spectroscopy after 10 h of carbonation (Figures S8 and S9, respectively). The Raman spectrum of MgO-2.5K reveals a broad peak centered in the range 1040–1130 cm–1, while MgO-5K exhibited a broad peak in the range 1040–1140 cm–1. By applying Gaussian deconvolution to the Raman spectra, we identified that the broad peak in both MgO-5K and MgO-2.5K can be deconvoluted into three carbonate phases with their peaks located at 1061 cm–1, 1082 cm–1 and 1103 cm–1 whereby MgO-5K contains a significantly larger fraction of the carbonate phase that has a characteristic peak at 1103 cm–1. Measurements of several reference carbonates (Figure S7) revealed that the peaks at 1061 cm–1 and 1082 cm–1 can be assigned to K2CO3 and baylissite (K2Mg(CO3)2·4H2O), respectively. The peak at 1103 cm–1 is most likely due to nesquehonite (MgCO3·3H2O), although the peak was slightly shifted compared to the reference position (1101 cm–1). This shift could be explained by the partial dehydration of the nesquehonite-type structure during carbonation at 315 °C, as nesquehonite dehydrates at temperatures above 270 °C. , Indeed, Hales et al. observed that when heating nesquehonite to 400 °C in air, the Raman signal of the main carbonate peak gradually shifted from 1098 cm–1 to 1105 cm–1, associated with the formation of an amorphous magnesium carbonate phase as a result of nesquehonite dehydration. , Hence, Raman spectroscopy combined with XRD analysis indicates the presence of amorphous carbonates in MgO-2.5K and MgO-5K that have been exposed for 10 h to carbonation conditions, whereby the amorphous carbonate phase exhibits a variety of local motifs that are similar to those found in baylissite (K2Mg(CO3)2·4H2O), K2CO3 and nesquehonite (MgCO3·3H2O).
In Situ XRD and Raman Analysis of MgO-5K
To investigate the evolution of the different carbonates of the most active sorbent, i.e. MgO-5K during carbonation at 315 °C in CO2, we conducted in situ XRD and in situ Raman spectroscopy experiments (Figure ). In the in situ XRD experiment (Figure a), MgO-5K was first pretreated at 450 °C in N2 for 1 h to remove adsorbed H2O and CO2, and to decompose hydrated phases such as baylissite (K2Mg(CO3)2·4H2O) and K2CO3·1.5H2O (Figure S10). After such pretreatment, the sorbent was cooled down to 315 °C in N2. The XRD pattern collected at 315 °C in N2, i.e. just before carbonation, revealed only Bragg peaks due to MgO and α-K2CO3. α-K2CO3 (hexagonal) is the high-temperature polymorph of K2CO3, which is stable at temperatures above 420 °C. , Upon cooling down the sorbent to 315 °C, a phase transition to β-K2CO3 (monoclinic) would typically be expected. However, only a minor fraction of α-K2CO3 converted into the β-phase, possibly due to kinetic limitations, and most of the K2CO3 remained as α-K2CO3 at the onset of the carbonation (Figure S11). To initiate carbonation, the gas atmosphere was changed from pure N2 to an 80% CO2/20% N2 mixture. During the first 20 min of carbonation, the Bragg peaks corresponding to α-K2CO3 disappeared, and a broad, diffuse X-ray scattering background appeared, indicating the formation of an amorphous phase. After ca. 1 h of carbonation, a mixture of different crystalline carbonate phases started to appear, including baylissite (K2Mg(CO3)2·4H2O), nesquehonite (MgCO3·3H2O), β-K2CO3, γ-K2CO3, and a very small amount of K2Mg(CO3)2. The intensity of these phases gradually increased with increasing carbonation time, revealing a slow crystallization process. Note, however, that the amorphous signal remained present, but slightly reduced in intensity.
2.
(a) In situ XRD data acquired during the carbonation of MgO-5K at 315 °C in CO2, overlaid with reference diffractograms of nesquehonite (MgCO3·3H2O), baylissite (K2Mg(CO3)2·4H2O), K2Mg(CO3)2, β-K2CO3, γ-K2CO3, and MgO. The dashed lines indicate the position of the background signal. The slight mismatch between the reference patterns and the experimental data arises from thermal lattices expansion at 315 °C. (b) In situ Raman spectroscopy data in the symmetric stretching region of carbonates, recorded after 0, 0.5, 1, 2, and 5 h of carbonation of MgO-5K at 315 °C in CO2. Deconvolution using Gaussian functions reveals peaks corresponding to K2CO3, baylissite, and nesquehonite. (c) Area under the v1 Raman peaks of K2CO3, baylissite, and nesquehonite as a function of carbonation time, derived from the deconvolution results in (b). (d–f) Quantitative analyses of the in situ XRD data shown in (a): (d) MgO conversion, (e) the evolution of the scattering signal due to amorphous phase(s), and (f) the evolution of the crystalline carbonates as a function of the carbonation time.
To complement the in situ XRD results and gain more insight into the evolution of the amorphous phase, we performed in situ Raman spectroscopy (Figure b (carbonate region) and Figure S12 (full spectral range)). As in the in situ XRD experiment, also in the in situ Raman spectroscopy experiments the sorbent MgO-5K was pretreated at 450 °C in N2 for 30 min, subsequently cooled down to 315 °C in N2 and exposed to CO2 at 315 °C for 5 h. At the start of the carbonation reaction, a single Raman band located at 1057 cm–1 was observed, corresponding to the ν1 symmetric stretching mode of K2CO3, confirming the absence of other carbonate phases. This v1 symmetric stretching mode of K2CO3 was shifted to a lower wavenumber compared to its room-temperature reference due to the thermal expansion of the K2CO3 lattice. During the first hour of carbonation, this carbonate band gradually shifted from 1057 cm–1 to 1077 cm–1. Gaussian peak fitting analysis of this band after 30 min and 1 h of carbonation revealed that this signal was a convolution of two contributions, viz. 1080 cm–1 (baylissite (K2Mg(CO3)2·4H2O)) and 1057 cm–1 (K2CO3), with the amount of the baylissite-type structure increasing over time. As carbonation proceeded, the carbonate peak shifted further to higher wavenumbers, reaching 1104 cm–1 while maintaining a shoulder at ca. 1057 cm–1. Gaussian fitting of the carbonate peak after 2 and 5 h of carbonation suggests that the peaks contained now three components: 1057 cm–1 (K2CO3), 1080 cm–1 (baylissite (K2Mg(CO3)2·4H2O)) and 1104 cm–1 (nesquehonite (MgCO3·3H2O)), whereby the nesquehonite-type structure is the main component. In reaction stage I (after 30 min and 1 h), peak broadening of the ν1 carbonate band was observed, consistent with the formation of disordered, amorphous phases, as confirmed by in situ XRD analysis (Figure a). In reaction stage II (after 2 and 5 h), the ν1 carbonate band narrowed, reflecting a partial crystallization of the carbonates. It is worth noting that the in situ Raman results are in line with ex situ Raman measurements conducted on different areas of MgO-5K after 30 min and 10 h of carbonation (Figure S9). However, after 30 min of carbonation the ex situ sample no longer shows unreacted K2CO3. This observation could be explained by a higher MgO conversion in the TGA compared to the Raman reaction cell, or due to a further reaction upon exposure to air, since baylissite can also form at room temperature by reaction with atmospheric CO2.
By tracking the intensity of the Raman bands during carbonation (Figure c), we find that in reaction stage I, the amount of K2CO3 decreased while there was a simultaneous emergence of a baylissite-type structure (a mixed K–Mg carbonate), indicating that the baylissite-type structure formed through the reaction of K2CO3 with MgO and CO2. During reaction stage II, a nesquehonite-type structure (i.e., a Mg-rich carbonate) appeared and the amount of it gradually increased. Note that in this stage, the amount of K2CO3 remained stable, while the amount of the baylissite-type structure decreased. This suggests that the baylissite-type structure evolved into a nesquehonite-type structure by incorporating additional Mg and CO2. At the end of the carbonation reaction (i.e., after 5 h), the material consisted predominantly of nesquehonite-type domains.
A quantitative analysis of the in situ XRD data provided further insight into the carbonation mechanism (see methods for details). Specifically, the MgO conversion (Figure d), the evolution of the amorphous signal (Figure e), and the evolution of the amount of the various crystalline phases (Figure f) were determined from the in situ XRD data. The conversion of MgO in MgO-5K during reaction stage I was determined as 12.0 ± 2.0% and increased to 16.0 ± 1.5% after 5 h of carbonation (Figure d). Note that the conversion of MgO after 10 h of carbonation, calculated based on the TGA results, was only 12% (determined using Eq. S3). The slightly higher conversion of MgO in the in situ XRD is likely due to differences in the gas–solid contact pattern. The evolution of the amorphous signal as a function of the carbonation time (Figure e) indicates that a large amount of the amorphous phase rapidly formed in reaction stage I, followed by a slight decrease in the total amount of the amorphous phase in reaction stage II. Tracking the evolution of the amount of crystalline carbonates (β-K2CO3, baylissite (K2Mg(CO3)2·4H2O) and nesquehonite (MgCO3·3H2O)) (Figure f), reveals that the different carbonates appeared simultaneously and that their appearance matches well with the onset of the decrease of the amount of the amorphous phase. Quantification of the in situ XRD data also indicates that the main fraction of the carbonates formed exists in an amorphous state with only a small fraction being in the form of crystalline carbonates.
To analyze the effect of promoter loading on the observed phase transformations, we conducted an in situ XRD measurement during the carbonation of MgO-7.5K (Figure S13). The phase transformations closely resembled those observed in MgO-5K. However, in reaction stage I, a fraction of the K2CO3 remained unreacted and did not become amorphous, consistent with the TGA results (Figure e and Table S2). Furthermore, after 5 h of carbonation, MgO-7.5K contained significantly more K2Mg(CO3)2 (Figure S14), which can be attributed to the higher K2CO3 loading, in combination with a comparable surface hydroxide concentration to MgO-5K (the surface areas of MgO-5K (25 m2 g–1) and MgO-7.5K (21 m2 g–1) are similar, see Figure S6). Hence, the overall water content in the amorphous double carbonate formed in reaction stage I is lower in MgO-7.5K compared to MgO-5K. The slower reaction kinetics observed in reaction stage II for MgO-7.5K compared to MgO-5K, suggest that the lower water content in the amorphous phase affects the further CO2 uptake.
The Role of Surface Hydroxides in the CO2 Uptake Mechanism
To investigate the origin of the water required to form hydrated carbonates (baylissite (K2Mg(CO3)2·4H2O) and nesquehonite (MgCO3·3H2O)) during the carbonation of MgO-5K at 315 °C in CO2, in situ DRIFTS experiments were conducted (Figure ). Specifically, DRIFTS spectra of MgO-5K were collected after pretreatment at 450 °C in N2 for 30 min and after carbonation at 315 °C in CO2 for 1 h, Figure a (hydroxide region) and Figure S16b (carbonate region). The DRIFTS spectrum collected at 315 °C in N2 (following the pretreatment and prior to carbonation) (Figure a), showed a sharp peak at 3740 cm–1 corresponding to isolated −OH groups, and a broad band in the 3700–3200 cm–1 region, which contains several features at 3620, 3513, and 3374 cm–1. The broad band can be assigned either to adsorbed water or multicoordinated H-bonded −OH groups, which can recombine to form water at temperatures below 400 °C. − After 1 h of carbonation, the content of surface hydroxides on MgO-5K remained largely unchanged, indicating that no water was captured nor released during carbonation, as also confirmed by TGA-MS experiments (Figure S15). Interestingly, the small amount of isolated −OH groups present on the surface of MgO in MgO-5K after pretreatment almost completely disappeared after carbonation, and only the broad band in the 3700- 3200 cm–1 region, corresponding to adsorbed water or crystal water, remained. The hydroxyl groups in the 3700–3200 cm–1 region can be assigned to the vibrational modes of −OH groups and H2O molecules in the crystal structure of baylissite and nesquehonite. Note that after pretreatment, the type of surface hydroxides in unpromoted MgO (MgO-0K, Figure S17) was similar to that in MgO-5K. However, there was a larger fraction of isolated −OH groups (peak at 3740 cm–1) in MgO-0K. Furthermore, unlike in MgO-5K, such isolated −OH groups remained on the surface of MgO after the carbonation of MgO-0K.
3.

DRIFTS spectra of the hydroxide region of MgO-5K after pretreatment in N2 for 30 min at (a) 450 °C, (b) 600 °C, and (c) 720 °C. The darker curves represent the spectra obtained at 315 °C in N2 after pretreatment, while the lighter curves correspond to the spectra obtained after an additional carbonation step at 315 °C in CO2 for 1 h. (d) CO2 uptake curves obtained in a TGA after pretreatment at 450 °C (yellow), 600 °C (black), and 750 °C (green).
To examine the effect of the pretreatment temperature on the amount of surface hydroxides in MgO-5K, in situ DRIFTS experiments were performed for pretreatment temperatures of 600 °C (Figure b) and 720 °C (Figure c). The DRIFTS spectrum of MgO-5K after pretreatment at 600 °C in N2 for 30 min (denoted as MgO-5K-pre600) showed a lower intensity of the broad −OH band compared to the spectrum after pretreatment at 450 °C, confirming that a pretreatment at a higher temperature removed surface hydroxides more effectively. However, a significant amount of isolated −OH groups and H-bonded −OH groups remained. In contrast, pretreatment at 720 °C (MgO-5K-pre720) resulted in the near-complete removal of the isolated −OH groups and a substantial reduction in the amount of H-bonded −OH groups. Interestingly, during the carbonation of both MgO-5K-pre600 and MgO-5K-pre720, the peak due to isolated −OH groups almost completely disappeared. However, the intensity of the broad −OH band increased slightly, revealing that the sorbents adsorbed small quantities of residual water present in the gas stream or the surrounding environment of the in situ DRIFTS cell. These findings indicate that H-bonded −OH groups on the surface of K2CO3-promoted MgO cannot be avoided under typical CO2 capture conditions, even when applying high temperature pretreatments.
To better understand the role of water/surface hydroxides on the CO2 uptake and the underlying carbonation mechanism of K2CO3-promoted MgO, the CO2 uptake curves of MgO-5K pretreated at 450 °C, 600 or 750 °C were compared (Figure d). All sorbents exhibited a two-step CO2 uptake, but the CO2 uptake in both reaction stages decreased with increasing pretreatment temperature, viz. 0.099, 0.087, and 0.065 gCO2 gsorbent –1 for pretreatment temperatures of 450 °C, 600 and 750 °C, respectively. As we observe the two-step CO2 uptake behavior for all sorbents, there is no indication of a change in the carbonation mechanism for lower surface hydroxide coverages (viz. higher pretreatment temperatures). Furthermore, the reduction in CO2 uptake with increasing pretreatment temperature (a 34% reduction at 750 °C compared to 450 °C) does not scale with the much larger decrease in the quantity of surface hydroxides (a ca. 76% reduction at 720 °C compared to 450 °C, based on the integrated area in the 3200–3850 cm–1 region). Instead, the lower CO2 uptakes when using higher pretreatment temperatures (i.e., 600 and 750 °C) are most likely due to sintering. Indeed, the average crystallite size of MgO in MgO-5K after carbonation increased with increasing pretreatment temperature, viz. 20 nm (450 °C), 24 nm (600 °C) and 34 nm (750 °C) (Figure S18), indicative of a higher degree of sintering with increasing temperature. Overall, our results suggest that variations in the amount of water/surface hydroxides do not alter the carbonation mechanism. Instead, the formation of crystalline hydrated phases such as nesquehonite and baylissite occurs due to the unavoidable presence of surface hydroxides under typical carbonation conditions.
Finally, the role of water in the carbonation mechanism was investigated by carbonating MgO-5K in humid CO2 (Figure S19). Compared to dry CO2, carbonation under humid conditions resulted in a 32% lower CO2 uptake, with only one CO2 uptake step being observed. Interestingly, when switching from humid to dry CO2, the CO2 uptake recovers rapidly, and the second CO2 uptake step appears. To probe these effects, DRIFTS spectra were obtained during carbonation in humid CO2. After 1 h of carbonation in humid CO2, the spectrum (Figure S20) closely resembles that obtained after carbonation in dry CO2 (Figure a and Figure S16b), i.e. there is no evidence of bicarbonates, indicating that humidity does not affect the nature of the adsorbed carbonates. Upon switching from humid to dry CO2, a slight decrease in the amount of adsorbed water (3200–3700 cm–1) was observed, with no detectable changes in the surface carbonates. This suggests that only bulk carbonates, to which DRIFTS is less sensitive, were affected. Furthermore, based on TGA results, only a small additional CO2 uptake was observed after switching to dry CO2 and holding for one hour (Figure S19). Finally, the carbonation kinetics may differ in the DRIFTS setup.
To further rationalize these findings, density functional theory (DFT) calculations were performed. CO2 adsorption energies on baylissite (K2Mg(CO3)2·4H2O) and partially dehydrated and fully dehydrated baylissite surfaces were calculated as −0.46 eV and −0.57 eV, respectively (Table S4 and Figure S21), indicating that CO2 adsorption is energetically more favorable onto a fully dehydrated baylissite surface as compared to a partially dehydrated one. These results suggest that humidity affects the CO2 uptake of K2CO3-promoted MgO, potentially due to stabilization of the baylissite-type amorphous phase and/or a competition between CO2 and H2O for the available sorption sites. This observation aligns with the well-established role of water in mineralization pathways, such as the stabilization of metastable amorphous intermediates. ,
Carbonation Mechanism
Combining insights from the analyses presented above, we propose a mechanism for the carbonation of K2CO3-promoted MgO (schematically shown in Figure ). In reaction stage I, a K-rich amorphous phase forms rapidly, which consists predominantly of baylissite-type (K–Mg carbonate) domains as well as some K2CO3-type domains (unreacted K2CO3). More specifically, the amorphous K–Mg carbonate forms through the reaction of K2CO3 with MgO, CO2, and residual water that is adsorbed on the surface of MgO via the following equation:
| 3 |
4.
Schematic representation of the evolution of the different crystalline and amorphous phases during the carbonation of K2CO3-promoted MgO.
In reaction stage I, x ≤ 1 as the amorphous carbonate is rich in K2CO3.
The rapid kinetics of double carbonate formation via eq can be rationalized by considering the binding strength of CO2 in baylissite (K2Mg(CO3)2·4H2O). DFT calculations of MgCO3, baylissite and partially dehydrated baylissite yield CO2binding strengths of −1.50 eV, −4.7 eV and −1.91 eV, respectively (Table S4). These values indicate that the binding strength of CO2 is much higher in baylissite (and partially dehydrated baylissite) than in MgCO3.
In reaction stage II, the K-rich amorphous carbonate further reacts with additional MgO and CO2, forming a Mg-rich amorphous carbonate (x > 1 in eq ), that mostly contains nesquehonite-type domains. Concurrently, small amounts of crystalline carbonates - including K2CO3, nesquehonite (MgCO3·3H2O) and baylissite (K2Mg(CO3)2·4H2O) - appear.
The promotional mechanism differs from that reported for Na2CO3 and CaCO3, in which crystalline double carbonates (NaMg(CO3)2 and CaMg(CO3)2) form rapidly and likely act as nucleation seeds for MgCO3 in the presence of molten nitrates. ,, In the case of K2CO3, a hydrated amorphous double carbonate forms that, however, does not act as a nucleation seed for MgCO3 formation. Instead, we hypothesize that this amorphous phase promotes carbonation in a manner analogous to molten alkali metal nitrates such as NaNO3, i.e. by dissolving MgO and CO2 and thereby facilitating the formation of a Mg-rich amorphous carbonate – even in the absence of a molten alkali metal nitrate salt.
Cyclic CO2 Uptake Performance of MgO-5K
To evaluate the cyclic CO2 uptake and release performance of K2CO3-promoted MgO, MgO-5K was pretreated at 450 °C in N2 for 30 min, carbonated for 5 h at 315 °C in CO2, and regenerated at 450 °C in N2 for 15 min in a TGA. The carbonation-regeneration cycle was repeated five times. During the first five carbonation-regeneration cycles (Figure a), the CO2 uptake decreased continuously from 0.114 gCO2 gsorbent –1 to 0.058 gCO2 gsorbent –1 (i.e., a reduction by 49%). While there was a slight decrease in the CO2 uptake in reaction stage I; most of the decrease in the CO2 uptake occurred in reaction stage II due to an extended induction period before its onset. HAADF-STEM images combined with elemental mapping of carbonated MgO-5K after the first (Figure b,c) and the fifth (Figure d,e) cycle revealed that while potassium was initially distributed uniformly, it partially agglomerated after five cycles. The increasingly heterogeneous distribution of potassium, and consequently of the K-rich intermediate with cycle number resulted in a decrease in the interface area with MgO. However, a high interfacial surface area between the K-rich amorphous carbonate and MgO is crucial to allow for a high additional CO2 uptake in reaction stage II. The XRD pattern of the carbonated sorbent in the fifth cycle (Figure S22) confirms that after 5 h of carbonation, no crystalline carbonates were detected. Instead, only an amorphous halo, most likely due to the baylissite-type phase, was present, consistent with TGA data indicating that reaction stage II had just started. These results suggest that while MgO-5K is suitable for cyclic CO2 capture from a process perspective, the segregation of K2CO3 significantly reduces its CO2 uptake capacity with cycle number. Therefore, future research should focus on maintaining a uniform K2CO3 distribution within the sorbent, e.g. by nanostructuring of the sorbent or by the addition of inert stabilizers, such as Al2O3, SiO2 or ZrO2. Such stabilizers have been studied extensively for CaO-based CO2 sorbents where they are shown to reduce sorbent sintering by acting as physical barriers between adjacent sorbent grains, thereby stabilizing the pore network of the sorbent. −
5.
(a) Cyclic CO2 uptake of MgO-5K over five consecutive carbonation–regeneration cycles recorded in a TGA (carbonation: 5 h at 315 °C in CO2; regeneration: 15 min at 450 °C in N2). (b) Plot of the temperature profile used during the TGA experiment in (a). HAADF-STEM images and elemental mapping (Mg and K) of carbonated MgO-5K (c, d) after the first cycle and (e, f) after five carbonation-regeneration cycles.
Conclusions
MgO-based sorbents are cost-efficient materials for CO2 capture, yet their slow kinetics require the use of promoters, typically alkali metal salts. Here, we elucidated the mechanism by which K2CO3 promotes the CO2 uptake of MgO and identified the carbonate phases that form during carbonation at 315 °C in CO2, using in situ XRD and in situ Raman spectroscopy. Initially, there was a rapid (<1 min) formation of a K-rich amorphous carbonate (K2CO3·xMgCO3·yH2O); in a second CO2 uptake step this intermediate phase further reacted with MgO and CO2, resulting in the formation of a Mg-rich amorphous carbonate. Furthermore, small amounts of hydrated crystalline carbonates, including nesquehonite (MgCO3·3H2O) and baylissite (K2Mg(CO3)2·4H2O), formed. In situ DRIFTS revealed that the water required for the formation of the hydrated carbonate phases was present in the form of surface hydroxyl groups at the onset of the carbonation reaction, even after a thermal pretreatment at 720 °C in N2. Introducing steam during carbonation, resulted in a lower CO2 uptake and inhibited the second CO2 uptake step, most likely by stabilizing the amorphous baylissite-type intermediate. It was also demonstrated that the sorbent can be used for cyclic CO2 capture operation, although minimizing K2CO3 segregation will be crucial to maintain a satisfying long-term cyclic performance. Overall, this study identifies a K-rich amorphous carbonate as the key promoter for CO2 uptake in K2CO3-promoted MgO and provides valuable insights to guide the design of more effective, nitrate-free MgO-based sorbents for CO2 capture. Furthermore, the phases identified here are likely to play also an important role in MgO-based sorbents that are promoted with a mixture of Na and K nitrates and carbonates, which are currently the best performing MgO-based CO2 sorbents reported in the literature.
Experimental Section
Chemicals
All chemicals were used as received: Magnesium hydroxide (Mg(OH)2, SLR, Fisher Chemical), potassium carbonate (K2CO3, extra pure, Fisher Chemical), sodium carbonate (Na2CO3, 99.5%, Acros Organics), magnesium nitrate hexahydrate (Mg(NO3)2·6H2O, Thermo Scientific), potassium nitrate (KNO3, ≥ 99.0%, Merck), hydromagnesite (Mg5(CO3)4(OH)2·4H2O, extra pure, Acros Organics), magnesium chloride hexahydrate (MgCl2·6H2O, ≥ 99.0%, Sigma-Aldrich) and ethylene glycol (>99%, Sigma-Aldrich).
Materials Preparation
K2CO3-Promoted MgO Sorbents
MgO was prepared by calcining Mg(OH)2 for 5 h at 500 °C in air. For the preparation of MgO-5K2CO3 (MgO promoted with 5 mol % K2CO3), 1 g of MgO, 0.17 g of K2CO3 and 10 mL isopropanol were loaded into a Pulverisette 7 planetary micro mill (Fritsch) at 150 rpm for 20 h (60 cycles of 10 min, followed by a 10 min break after each cycle) using 1 mm ZrO2 balls in a Si3N4 cylinder. The ball-milled material was collected and dried in an oven at 100 °C for 2 h. The same method was used to prepare MgO-2.5K2CO3, MgO-7.5K2CO3 and the unpromoted MgO reference by adjusting the amount of K2CO3. Finally, the sorbents were calcined at 500 °C for 4 h.
Reference Materials
Baylissite (K2Mg(CO3)2·4H2O)
was synthesized using an adapted method from literature. Specifically, 1 mL of a magnesium nitrate solution was added dropwise to a concentrated K2CO3 solution (3 g K2CO3 in 10 mL H2O) at 90 °C under constant stirring. After 10 min, the stirring was stopped, and the temperature was lowered to 60 °C. The mixture was maintained at 60 °C for 72 h to allow crystallization. The resulting mixture was then filtered (without washing as washing would have dissolved the crystals) and dried at room temperature. XRD, Raman and SEM-EDX analysis (Figures S3 and S4), revealed that the synthesized baylissite was almost phase-pure with only a small K2CO3 impurity.
K2Mg(CO3)2
was prepared by mixing as-synthesized baylissite with 20 mol % KNO3 upon grinding. The material was subsequently treated in a TGA at 340 °C for 3 h in CO2 (80 mL/min and a heating rate of 50 °C/min) to remove the crystal water, resulting in KNO3-promoted K2Mg(CO3)2 (see Figure S23 for more details).
MgCO3
was synthesized from hydromagnesite via a wet chemistry method, using a suspension of hydromagnesite in a 95 wt % ethylene glycol solution in water. This suspension was refluxed at 150 °C under a continuous flow of CO2. The phase purity was verified by XRD (Figure S24).
Nesquehonite (MgCO3·3H2O)
was synthesized using a method adopted from the literature. Solutions of 0.5 M MgCl2 and 0.5 M Na2CO3 were mixed at a volume ratio of 1:1 and stirred at 1000 rpm for 2 h at room temperature. The resulting white suspension was filtered out, washed with distilled water, collected and dried overnight in air. The phase purity was verified by XRD (Figure S25a).
Nesquehonite-Type Amorphous Magnesium Carbonate
was prepared by heating nesquehonite to 315 °C in a TGA under a N2 atmosphere at a heating rate of 5 °C min–1, resulting in a weight loss of 36 wt %. This value aligns closely with the theoretical weight loss for the dehydration of MgCO3·3H2O (39 wt %), indicating the nearly complete removal of crystal water (Figure S25b). The amorphous nature of the resulting magnesium carbonate was confirmed by the absence of crystalline peaks in the XRD pattern (Figure S25a).
Characterization Methods
Raman Spectroscopy
Raman spectra were collected with a Thermo Scientific DXR2 Raman spectrometer equipped with a 532 nm laser using a spot size of 1.8 μm. The spectra were acquired in the range of 100–3,500 cm–1 with a spectral resolution of 0.964 cm–1. For the ex situ measurements, the Raman spectra represent the average obtained from measurements taken in five different regions of each sample. For the in situ measurements, 20 mg sample was loaded into a Linkam CCR1000 in situ cell. The sample was subjected to a pretreatment for 30 min at 450 °C in N2, followed by carbonation for 5 h at 315 °C in CO2. The heating and cooling rates were set to 30 °C min–1 and a gas flow rate of 30 mL min–1 was used. The Raman band intensity was obtained from the in situ Raman data by integrating the area under the ν1 bands for K2CO3, baylissite and nesquehonite. The areas were obtained through deconvolution using Gaussian functions (Figure b) as a function of the carbonation time. Note that the deconvolution results are inherently semiquantitative: while they allowed us to track the temporal changes in the area under each individual Raman band, they do not enable a reliable quantification of the relative amounts of the different phases. Additionally, the Raman bands reflect distinct local structural motifs within the amorphous phase rather than specific crystalline phases. These motifs are analogous to those observed in the reference carbonates used for the deconvolution.
CO2 Uptake Measurements
TGA experiments were carried out in a Mettler Toledo TGA/DSC 3+. In a typical analysis, 15 mg of sample (or 50 mg when a mass spectrometer was connected for product gas analysis) was loaded into a 70 μL alumina crucible. For a typical carbonation experiment, the flow of reactive gas (N2 for pretreatment and regeneration and CO2 for carbonation) was set to 80 mL min–1 and a heating rate of 50 °C min–1 was used. To introduce moisture (humidity) into the system, the reactive gas was sent through a gas washing bottle at room temperature to achieve a relative humidity of close to 100% at room temperature (which corresponds to a relative humidity of ca. 4.5% at 315 °C). For selected experiments, the gas composition at the outlet of the TGA was analyzed with a mass spectrometer (MKS Cirrus TM 3-XD), focusing on the signal for the following masses (m/z): 18 (H2O), 28 (N2), 32 (O2) and 44 (CO2).
X-ray Diffraction
XRD patterns were collected using a PANalytical Empyrean X-ray powder diffractometer (45 kV and 40 mA using Cu Kα radiation) equipped with a Bragg–Brentano HD mirror. Ex situ scans were collected in the 2θ range of 8–70 ° (step size 0.033 ° and time per step 3.2 s). For the in situ measurements, 20 mg sample was placed onto a quartz disc and loaded into an Anton Paar XRK 900 reaction chamber. XRD patterns were collected in the 2θ range of 8–50° (step size of 0.12 ° and time per step of 3.4 s), while the sample was subjected to pretreatment (at 450 °C in N2) and carbonation (at 315 °C in 80 vol% CO2/N2) conditions using a gas flow rate of 200 mL min–1 and a heating rate of 20 °C min–1.
Phase Quantifications Using the in Situ XRD Data
The MgO conversion was calculated as (AMgO(200)(t = 0) – AMgO(200)(t)) × AMgO(200)(0)−1 × 100%, where AMgO(200)(t) is the area under the (200) peak of MgO (2θ = 42.9°) at a given time t. It should be noted that this calculation considers crystalline phases only. To estimate the amount of the amorphous phase as a function of carbonation time, the background of the XRD pattern at carbonation t = 0 (determined using a linear interpolation) was subtracted from all subsequent patterns collected during carbonation. The amorphous fraction was then calculated by fitting the remaining (adjusted) background using a linear interpolation (spline) and integrating the area under the interpolation. The evolution of the crystalline phases was determined by integrating the area under the most intense, nonoverlapping Bragg peak for each phase as a function of carbonation time: α-K2CO3 (102) (2θ = 31.1°), β-K2CO3 (130) (2θ = 31.4°), baylissite (120) (2θ = 29.7°) and nesquehonite (200) (2θ = 23.1°). Note that the results for both the amorphous and the crystalline phases are semiquantitative. While they provide insight into the temporal evolution of the amount of each phase, they do not allow for a direct comparison between the different phases.
Electron Microscopy
Transmission electron microscopy (TEM) and scanning TEM combined with energy dispersive X-ray spectroscopy (STEM-EDX) measurements were acquired with a FEI Talos F200X electron microscope operated at 200 kV and equipped with a Super-X EDS system. SEM-EDX measurements were obtained using a Quanta 200F SEM microscope operated at 10 kV. Potassium loadings based on the SEM-EDX results were determined by averaging the quantification results from 3 different regions of the sample.
N2 Physisorption
BET surface areas of the materials were determined from a N2 physisorption isotherm recorded at 77 K on an Anton Paar Nova 800 apparatus. The samples were degassed at 350 °C under vacuum (10–3 mbar) for 3 h prior to measurement.
ICP-OES
The relative amounts of Mg and K in the sorbents were determined using an Agilent 5100 VDV ICP-OES instrument. Samples were dissolved in aqua regia prior to the measurements and the measurements were performed in triplicate for reproducibility.
In Situ DRIFTS
In situ DRIFTS experiments were performed using a Nicolet 6700 FTIR spectrometer (resolution 4 cm–1) coupled with a Harrick Praying Mantis high temperature reaction chamber with ZnSe windows. The sample holder was filled with quartz wool and ca. 10 mg of ground sample was placed on top and pressed flat. In a typical experiment, the sample was heated to the pretreatment temperature (e.g., 450 °C, 600 or 720 °C) in N2 and kept at this temperature for 30 min, followed by cooling down to 315 °C. At 315 °C, the sample was exposed to CO2 for 2 h. The heating rate was set to 10 °C min–1 and the gas flow rate was 30 mL min–1. The FTIR signal of KBr powder treated under similar conditions was used as a spectral background.
DFT Calculations
DFT calculations were performed using the Projector Augmented Wave (PAW) method , as implemented in the Vienna Ab initio Simulation Package (VASP). , The calculations were completed with a plane-wave cutoff energy of 600 eV and Gamma k-points. The electronic self-consistent calculation converged to 1 × 10–6 eV and ionic relaxation steps were performed using the conjugate-gradient method (IBRION = 2) and continued until the total force on each atom was below a tolerance of 0.01 eV/Å. The generalized gradient approximation (GGA) was used for the exchange correlation functionals as parametrized by Perdew–Burke–Ernzerhof (PBE). The bulk and (100) surface calculations, including MgO, MgCO3 and baylissite, were conducted to study the adsorption/binding of CO2 in/on different systems, both bulk and surfaces with the presence and absence of water molecules. The binding of CO2 (Eb‑CO2) with the substrate was calculated using the following equation.
| 4 |
where E [substrate + CO2], E[substrate] and E[CO2]are the electronic energies of the adsorbed system, the clean bulk/surface, and CO2, respectively.
Supplementary Material
Acknowledgments
The authors are grateful to the Scientific Center for Optical and Electron Microscopy (ScopeM, ETH Zürich) for the use of their electron microscopy facilities. We also thank Dr. A. Kierzkowska for performing the ICP-OES measurements. Financial support by the Swiss National Science Foundation (SNSF, 200020_156015) is acknowledged.
The Supporting Information is available free of charge at . The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacsau.5c00838.
Additional description of methods, XRD patterns, and Raman spectra of the reference materials, calculation of MgO conversion, and supplementary results including ICP-OES, XRD, Raman, TGA(-MS), DRIFTS, SEM-EDX, and N2 physisorption data (PDF)
CRediT: Annelies Landuyt conceptualization, data curation, formal analysis, investigation, methodology, visualization, writing-original draft; Felix Donat conceptualization, investigation, supervision, validation, writing-review and editing Paula M. Abdala conceptualization, supervision, validation, writing-review and editing; Christoph R. Müller conceptualization, funding acquisition, project administration, resources, supervision, validation, writing-review and editing.
The authors declare no competing financial interest.
References
- Wilberforce T., Olabi A. G., Sayed E. T., Elsaid K., Abdelkareem M. A.. Progress in Carbon Capture Technologies. Sci. Total Environ. 2021;761:143203. doi: 10.1016/j.scitotenv.2020.143203. [DOI] [PubMed] [Google Scholar]
- Suleman, H. ; Fosbo̷l, P. L. ; Nasir, R. ; Ameen, M. . Sustainable Carbon Capture: Technologies and Applications; CRC Press: Boca Raton, 2022. [Google Scholar]
- MacDowell N., Florin N., Buchard A., Hallett J., Galindo A., Jackson G., Adjiman C. S., Williams C. K., Shah N., Fennell P.. An Overview of CO2 Capture Technologies. Energy Environ. Sci. 2010;3(11):1645. doi: 10.1039/c004106h. [DOI] [Google Scholar]
- Hu Y., Guo Y., Sun J., Li H., Liu W.. Progress in MgO Sorbents for Cyclic CO2 Capture: A Comprehensive Review. J. Mater. Chem. A. 2019;7(35):20103–20120. doi: 10.1039/C9TA06930E. [DOI] [Google Scholar]
- Dunstan M. T., Donat F., Bork A. H., Grey C. P., Müller C. R.. CO2 Capture at Medium to High Temperature Using Solid Oxide-Based Sorbents: Fundamental Aspects, Mechanistic Insights, and Recent Advances. Chem. Rev. 2021;121(20):12681–12745. doi: 10.1021/acs.chemrev.1c00100. [DOI] [PubMed] [Google Scholar]
- Wu C., Huang Q., Xu Z., Sipra A. T., Gao N., Vandenberghe L. P. de S., Vieira S., Soccol C. R., Zhao R., Deng S., Boetcher S. K. S., Lu S., Shi H., Zhao D., Xing Y., Chen Y., Zhu J., Feng D., Zhang Y., Deng L., Hu G., Webley P. A., Liang D., Ba Z., Mlonka-Mędrala A., Magdziarz A., Miskolczi N., Tomasek S., Lam S. S., Foong S. Y., Ng H. S., Jiang L., Yan X., Liu Y., Ji Y., Sun H., Zhang Y., Yang H., Zhang X., Sun M., Tsang D. C. W., Shang J., Muller C., Rekhtina M., Krödel M., Bork A. H., Donat F., Liu L., Jin X., Liu W., Saqline S., Wu X., Xu Y., Khan A. L., Ali Z., Lin H., Hu L., Huang J., Singh R., Wang K., He X., Dai Z., Yi S., Konist A., Baqain M. H. S., Zhao Y., Sun S., Chen G., Tu X., Weidenkaff A., Kawi S., Lim K. H., Song C., Yang Q., Zhao Z., Gao X., Jiang X., Ji H., Akinola T. E., Lawal A., Otitoju O. S., Wang M., Zhang G., Ma L., Sempuga B. C., Liu X., Oko E., Daramola M., Yu Z., Chen S., Kang G., Li Q., Gao L., Liu L., Zhou H.. A Comprehensive Review of Carbon Capture Science and Technologies. Carbon Capture Sci. Technol. 2024;11:100178. doi: 10.1016/j.ccst.2023.100178. [DOI] [Google Scholar]
- Gouedard C., Picq D., Launay F., Carrette P.-L.. Amine Degradation in CO2 Capture. I. A Review. Int. J. Greenh. Gas Control. 2012;10:244–270. doi: 10.1016/j.ijggc.2012.06.015. [DOI] [Google Scholar]
- Meng F., Meng Y., Ju T., Han S., Lin L., Jiang J.. Research Progress of Aqueous Amine Solution for CO2 Capture: A Review. Renew. Sustain. Energy Rev. 2022;168:112902. doi: 10.1016/j.rser.2022.112902. [DOI] [Google Scholar]
- Rochelle G. T.. Amine Scrubbing for CO2 Capture. Science. 2009;325(5948):1652–1654. doi: 10.1126/science.1176731. [DOI] [PubMed] [Google Scholar]
- Paltsev S., Morris J., Kheshgi H., Herzog H.. Hard-to-Abate Sectors: The Role of Industrial Carbon Capture and Storage (CCS) in Emission Mitigation. Appl. Energy. 2021;300:117322. doi: 10.1016/j.apenergy.2021.117322. [DOI] [Google Scholar]
- Najmi B., Bolland O., Colombo K. E.. A Systematic Approach to the Modeling and Simulation of a Sorption Enhanced Water Gas Shift (SEWGS) Process for CO2 Capture. Sep. Purif. Technol. 2016;157:80–92. doi: 10.1016/j.seppur.2015.11.013. [DOI] [Google Scholar]
- Choi D. H., Lee J. B., Eom T. H., Baek J. I., Jegarl S., Ryu C. K.. Study of MgO-Based Dry Regenerable Sorbent for Sorption Enhanced Water Gas Shift Reaction. Renew. Energy. 2013;54:144–149. doi: 10.1016/j.renene.2012.08.044. [DOI] [Google Scholar]
- Gregg S. J., Ramsay J. D.. Adsorption of Carbon Dioxide by Magnesia Studied by Use of Infrared and Isotherm Measurements. J. Chem. Soc. A Inorganic, Phys. Theor. 1970:2784–2787. doi: 10.1039/j19700002784. [DOI] [Google Scholar]
- Bhagiyalakshmi M., Lee J. Y., Jang H. T.. Synthesis of Mesoporous Magnesium Oxide: Its Application to CO2 Chemisorption. Int. J. Greenh. Gas Control. 2010;4(1):51–56. doi: 10.1016/j.ijggc.2009.08.001. [DOI] [Google Scholar]
- Gao W., Zhou T., Gao Y., Louis B., O’Hare D., Wang Q.. Molten Salts-Modified MgO-Based Adsorbents for Intermediate-Temperature CO2 Capture: A Review. J. Energy Chem. 2017;26(5):830–838. doi: 10.1016/j.jechem.2017.06.005. [DOI] [Google Scholar]
- Ruhaimi A. H., Aziz M. A. A., Jalil A. A.. Magnesium Oxide-Based Adsorbents for Carbon Dioxide Capture: Current Progress and Future Opportunities. J. CO2 Util. 2021;43:101357. doi: 10.1016/j.jcou.2020.101357. [DOI] [Google Scholar]
- Harada T., Simeon F., Hamad E. Z., Hatton T. A.. Alkali Metal Nitrate-Promoted High-Capacity MgO Adsorbents for Regenerable CO2 Capture at Moderate Temperatures. Chem. Mater. 2015;27(6):1943–1949. doi: 10.1021/cm503295g. [DOI] [Google Scholar]
- Dal Pozzo A., Armutlulu A., Rekhtina M., Abdala P. M., Müller C. R.. CO2 Uptake and Cyclic Stability of MgO-Based CO2 Sorbents Promoted with Alkali Metal Nitrates and Their Eutectic Mixtures. ACS Appl. Energy Mater. 2019;2(2):1295–1307. doi: 10.1021/acsaem.8b01852. [DOI] [Google Scholar]
- Cui H., Zhang Q., Hu Y., Peng C., Fang X., Cheng Z., Galvita V. V., Zhou Z.. Ultrafast and Stable CO2 Capture Using Alkali Metal Salt-Promoted MgO-CaCO3 Sorbents. ACS Appl. Mater. Interfaces. 2018;10(24):20611–20620. doi: 10.1021/acsami.8b05829. [DOI] [PubMed] [Google Scholar]
- Ding J., Yu C., Lu J., Wei X., Wang W., Pan G.. Enhanced CO2 Adsorption of MgO with Alkali Metal Nitrates and Carbonates. Appl. Energy. 2020;263:114681. doi: 10.1016/j.apenergy.2020.114681. [DOI] [Google Scholar]
- Rekhtina M., Krödel M., Wu Y.-H., Kierzkowska A., Donat F., Abdala P. M., Müller C. R.. Deciphering the Structural Dynamics in Molten Salt-Promoted MgO-Based CO2 Sorbents and Their Role in the CO2 Uptake. Sci. Adv. 2023;9(26):eadg5690. doi: 10.1126/sciadv.adg5690. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Landuyt A., Kochetygov I., McMonagle C. J., Kumar P. V., Yuwono J. A., Queen W. L., Abdala P. M., Müller C. R.. Role of Na2CO3 as Nucleation Seeds to Accelerate the CO2 Uptake Kinetics of MgO-Based Sorbents. JACS Au. 2024;4(12):4809–4820. doi: 10.1021/jacsau.4c00782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kruizenga A., Gill D.. Corrosion of Iron Stainless Steels in Molten Nitrate Salt. In Energy Procedia. Energy Procedia. 2014;49:878–887. doi: 10.1016/j.egypro.2014.03.095. [DOI] [Google Scholar]
- Jo S.-I., An Y.-I., Kim K.-Y., Choi S.-Y., Kwak J.-S., Oh K.-R., Kwon Y.-U.. Mechanisms of Absorption and Desorption of CO2 by Molten NaNO3-Promoted MgO. Phys. Chem. Chem. Phys. 2017;19(8):6224–6232. doi: 10.1039/C6CP07787K. [DOI] [PubMed] [Google Scholar]
- Cui H., Cheng Z., Zhou Z.. Unravelling the Role of Alkaline Earth Metal Carbonates in Intermediate Temperature CO2 Capture Using Alkali Metal Salt-Promoted MgO-Based Sorbents. J. Mater. Chem. A. 2020;8(35):18280–18291. doi: 10.1039/D0TA06170K. [DOI] [Google Scholar]
- Papalas T., Antzaras A. N., Lemonidou A. A.. Magnesite-Derived MgO Promoted with Molten Salts and Limestone as Highly-Efficient CO2 Sorbent. J. CO2 Util. 2021;53:101725. doi: 10.1016/j.jcou.2021.101725. [DOI] [Google Scholar]
- Papalas T., Polychronidis I., Antzaras A. N., Lemonidou A. A.. Enhancing the Intermediate-Temperature CO2 Capture Efficiency of Mineral MgO via Molten Alkali Nitrates and CaCO3 . Characterization and Sorption Mechanism. J. CO2 Util. 2021;50:101605. doi: 10.1016/j.jcou.2021.101605. [DOI] [Google Scholar]
- Ryu D. Y., Jo S., Kim T. Y., In S. Y., Kim J. K., Hwang J. E., Kim J. C., Lee S. C.. Influence of the Sorption Pressure and K2CO3 Loading of a MgO-Based Sorbent for Application to the SEWGS Process. Korean J. Chem. Eng. 2022;39(4):1028–1035. doi: 10.1007/s11814-021-0967-1. [DOI] [Google Scholar]
- Xiao G., Singh R., Chaffee A., Webley P.. Advanced Adsorbents Based on MgO and K2CO3 for Capture of CO2 at Elevated Temperatures. Int. J. Greenh. Gas Control. 2011;5(4):634–639. doi: 10.1016/j.ijggc.2011.04.002. [DOI] [Google Scholar]
- Lee S. C., Choi B. Y., Lee T. J., Ryu C. K., Ahn Y. S., Kim J. C.. CO2 Absorption and Regeneration of Alkali Metal-Based Solid Sorbents. Catal. Today. 2006;111(3–4):385–390. doi: 10.1016/j.cattod.2005.10.051. [DOI] [Google Scholar]
- Lee S. C., Kim J. C.. Dry Potassium-Based Sorbents for CO2 Capture. Catal. Surv. from Asia. 2007;11(4):171–185. doi: 10.1007/s10563-007-9035-z. [DOI] [Google Scholar]
- Lee S. C., Chae H. J., Lee S. J., Choi B. Y., Yi C. K., Lee J. B., Ryu C. K., Kim J. C.. Development of Regenerable MgO-Based Sorbent Promoted with K2CO3 for CO2 Capture at Low Temperatures. Environ. Sci. Technol. 2008;42(8):2736–2741. doi: 10.1021/es702693c. [DOI] [PubMed] [Google Scholar]
- Kwak J.-S., Kim K.-Y., Yoon J. W., Oh K.-R., Kwon Y.-U.. Interfacial Interactions Govern the Mechanisms of CO2 Absorption and Desorption on A2CO3-Promoted MgO (A = Na, K, Rb, and Cs) Absorbents. J. Phys. Chem. C. 2018;122(35):20289–20300. doi: 10.1021/acs.jpcc.8b04895. [DOI] [Google Scholar]
- Mayorga, S. G. ; Weigel, S. J. ; Gaffney, T. R. ; Brzozowski, J. R. . Carbon Dioxide Adsorbents Containing Magnesium Oxide Suitable for Use at High Temperatures, 2001. [Google Scholar]
- Duan Y., Zhang K., Li X. S., King D. L., Li B., Zhao L., Xiao Y.. Ab Initio Thermodynamic Study of the CO2 Capture Properties of M2CO3 (M = Na, K)- and CaCO3-Promoted MgO Sorbents Towards Forming Double Salts. Aerosol Air Qual. Res. 2014;14(2):470–479. doi: 10.4209/aaqr.2013.05.0178. [DOI] [Google Scholar]
- Lee J. M., Min Y. J., Lee K. B., Jeon S. G., Na J. G., Ryu H. J.. Enhancement of CO2 Sorption Uptake on Hydrotalcite by Impregnation with K2CO3 . Langmuir. 2010;26(24):18788–18797. doi: 10.1021/la102974s. [DOI] [PubMed] [Google Scholar]
- Dagie, R. ; Li, X. S. ; Spies, K. ; Rainbolt, J. ; Braunberger, B. ; King, D. ; Xing, R. ; Zhu, Y. ; Li, L. . Coal-Derived Warm Syngas Purification and CO2 Capture-Assisted Methane Production Final Report. PNNL-23777, Pacific Northwest Natl. Lab, 2014, 1076. [Google Scholar]
- Bucat R. B., Patrick J. M., White A. H., Willis A. C.. Crystal Structure of Baylissite, K2Mg(CO3)2.4H2O. Aust. J. Chem. 1977;30(6):1379–1382. doi: 10.1071/CH9771379. [DOI] [Google Scholar]
- Bork A. H., Rekhtina M., Willinger E., Castro-Fernández P., Drnec J., Abdala P. M., Müller C. R.. Peering into Buried Interfaces with X-Rays and Electrons to Unveil MgCO3 Formation during CO2 Capture in Molten Salt-Promoted MgO. Proc. Natl. Acad. Sci. U. S. A. 2021;118(26):e2103971118. doi: 10.1073/pnas.2103971118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ballirano P., De Vito C., Ferrini V., Mignardi S.. The Thermal Behaviour and Structural Stability of Nesquehonite, MgCO3·3H2O, Evaluated by in Situ Laboratory Parallel-Beam X-Ray Powder Diffraction: New Constraints on CO2 Sequestration within Minerals. J. Hazard. Mater. 2010;178(1–3):522–528. doi: 10.1016/j.jhazmat.2010.01.113. [DOI] [PubMed] [Google Scholar]
- Hales M. C., Frost R. L., Martens W. N.. Thermo-Raman Spectroscopy of Synthetic Nesquehonite Implication for the Geosequestration of Greenhouse Gases. J. Raman Spectrosc. 2008;39(9):1141–1149. doi: 10.1002/jrs.1950. [DOI] [Google Scholar]
- Åhlén M., Cheung O., Strømme M.. Amorphous Mesoporous Magnesium Carbonate as a Functional Support for UV-Blocking Semiconductor Nanoparticles for Cosmetic Applications. ACS Omega. 2019;4(2):4429–4436. doi: 10.1021/acsomega.8b03498. [DOI] [Google Scholar]
- Becht H. Y., Struikmans R.. A Monoclinic High-Temperature Modification of Potassium Carbonate. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1976;32(12):3344–3346. doi: 10.1107/S0567740876010303. [DOI] [Google Scholar]
- Schneider S. J., Levin E. M.. Polymorphism of K2CO3 . J. Am. Ceram. Soc. 1973;56(4):218–219. doi: 10.1111/j.1151-2916.1973.tb12461.x. [DOI] [Google Scholar]
- Frantz J. D.. Raman Spectra of Potassium Carbonate and Bicarbonate Aqueous Fluids at Elevated Temperatures and Pressures: Comparison with Theoretical Simulations. Chem. Geol. 1998;152(3–4):211–225. doi: 10.1016/S0009-2541(98)00058-8. [DOI] [Google Scholar]
- Wang D., Hamm L. M., Bodnar R. J., Dove P. M.. Raman Spectroscopic Characterization of the Magnesium Content in Amorphous Calcium Carbonates. J. Raman Spectrosc. 2012;43(4):543–548. doi: 10.1002/jrs.3057. [DOI] [Google Scholar]
- Knözinger E., Jacob K. H., Singh S., Hofmann P.. Hydroxyl Groups as IR Active Surface Probes on MgO Crystallites. Surf. Sci. 1993;290(3):388–402. doi: 10.1016/0039-6028(93)90721-U. [DOI] [Google Scholar]
- Liu P., Abdala P. M., Goubert G., Willinger M., Copéret C.. Ultrathin Single Crystalline MgO(111) Nanosheets**. Angew. Chemie Int. Ed. 2021;60(6):3254–3260. doi: 10.1002/anie.202013196. [DOI] [PubMed] [Google Scholar]
- Kumar A., Kumar J.. On the Synthesis and Optical Absorption Studies of Nano-Size Magnesium Oxide Powder. J. Phys. Chem. Solids. 2008;69(11):2764–2772. doi: 10.1016/j.jpcs.2008.06.143. [DOI] [Google Scholar]
- Skliros V., Tsakiridis P., Perraki M.. A Combined Raman, Fourier Transform Infrared, and X-ray Diffraction Study of Thermally Treated Nesquehonite. J. Raman Spectrosc. 2020;51(9):1445–1453. doi: 10.1002/jrs.5768. [DOI] [Google Scholar]
- Gardeh M. G., Kistanov A. A., Nguyen H., Manzano H., Cao W., Kinnunen P.. Exploring Mechanisms of Hydration and Carbonation of MgO and Mg(OH)2 in Reactive Magnesium Oxide-Based Cements. J. Phys. Chem. C. 2022;126(14):6196–6206. doi: 10.1021/acs.jpcc.1c10590. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prange M. P., Mergelsberg S. T., Kerisit S. N.. Structural Water in Amorphous Carbonate Minerals: Ab Initio Molecular Dynamics Simulations of X-Ray Pair Distribution Experiments. Phys. Chem. Chem. Phys. 2023;25(9):6768. doi: 10.1039/D2CP04881G. [DOI] [PubMed] [Google Scholar]
- Gao Y., Tao Y., Li G., Shen P., Pellenq R. J.-M., Poon C. S.. Moisture-Driven Carbonation Kinetics for Ultrafast CO2 Mineralization. Proc. Natl. Acad. Sci. U. S. A. 2025;122(1):e2418239121. doi: 10.1073/pnas.2418239121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Landuyt A., Kumar P. V., Yuwono J. A., Bork A. H., Donat F., Abdala P. M., Müller C. R.. Uncovering the CO2 Capture Mechanism of NaNO3-Promoted MgO by 18O Isotope Labeling. JACS Au. 2022;2(12):2731–2741. doi: 10.1021/jacsau.2c00461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim S. M., Liao W.-C., Kierzkowska A. M., Margossian T., Hosseini D., Yoon S., Broda M., Copéret C., Müller C. R.. In Situ XRD and Dynamic Nuclear Polarization Surface Enhanced NMR Spectroscopy Unravel the Deactivation Mechanism of CaO-Based, Ca3Al2O6-Stabilized CO2 Sorbents. Chem. Mater. 2018;30(4):1344–1352. doi: 10.1021/acs.chemmater.7b05034. [DOI] [Google Scholar]
- Yoon H. J., Lee K. B.. Introduction of Chemically Bonded Zirconium Oxide in CaO-Based High-Temperature CO2 Sorbents for Enhanced Cyclic Sorption. Chem. Eng. J. 2019;355:850–857. doi: 10.1016/j.cej.2018.08.148. [DOI] [Google Scholar]
- Krödel M., Landuyt A., Abdala P. M., Müller C. R.. Mechanistic Understanding of CaO-Based Sorbents for High-Temperature CO2 Capture: Advanced Characterization and Prospects. ChemSusChem. 2020;13(23):6259–6272. doi: 10.1002/cssc.202002078. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sandengen K., Jøsang L. O., Kaasa B.. Simple Method for Synthesis of Magnesite (MgCO3) Ind. Eng. Chem. Res. 2008;47(4):1002–1004. doi: 10.1021/ie0706360. [DOI] [Google Scholar]
- Yamamoto G., Kyono A., Okada S.. Temperature Dependence of Amorphous Magnesium Carbonate Structure Studied by PDF and XAFS Analyses. Sci. Rep. 2021;11(1):22876. doi: 10.1038/s41598-021-02261-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blöchl P. E.. Projector Augmented-Wave Method. Phys. Rev. B. 1994;50(24):17953–17979. doi: 10.1103/PhysRevB.50.17953. [DOI] [PubMed] [Google Scholar]
- Kresse G., Joubert D.. From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method. Phys. Rev. B. 1999;59(3):1758–1775. doi: 10.1103/PhysRevB.59.1758. [DOI] [Google Scholar]
- Kresse G., Furthmüller J.. Efficient Iterative Schemes for Ab Initio Total-Energy Calculations Using a Plane-Wave Basis Set. Phys. Rev. B. 1996;54(16):11169–11186. doi: 10.1103/PhysRevB.54.11169. [DOI] [PubMed] [Google Scholar]
- Kresse G., Furthmüller J.. Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors Using a Plane-Wave Basis Set. Comput. Mater. Sci. 1996;6(1):15–50. doi: 10.1016/0927-0256(96)00008-0. [DOI] [Google Scholar]
- Perdew J. P., Burke K., Ernzerhof M.. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996;77(18):3865–3868. doi: 10.1103/PhysRevLett.77.3865. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.




