Lotfi Boudjemaah,
Anil Kumar Dahiyab,
Ivan da Silva
c,
Diego Gianolio
d,
Izuchika Ndukab,
Manfred Erwin Schuster
e,
Gea Theodora van de Kerkhof
de,
Paulina Kalinowskab,
Emilio Borrego-Marin
f,
Jorge A. R. Navarro
f,
Valentina Colombo
i,
June McCorquodalegh,
David C. Grinter
d,
Pilar Ferrer
d,
Georg Held
d,
C. Richard A. Catlowagh and
Rosa Arrigo
*bd
aDepartment of Chemistry, University College London, 20 Gordon Street, London, WC1H 0AJ, UK
bSchool of Science, Engineering and Environment, University of Salford, Manchester, M5 4WT, UK. E-mail: r.arrigo@salford.ac.uk
cISIS Neutron and Muon Source, STFC Rutherford Appleton Laboratory, Didcot, Oxfordshire OX11 0QX, UK
dDiamond Light Source Ltd., Harwell Science & Innovation Campus, Didcot, Oxfordshire OX11 0DE, UK
eJohnson Matthey Technology Centre, Reading, RG4 9NH, UK
fDepartamento de Química Inorgánica, Universidad de Granada, Av. Fuentenueva S/N, 18071 Granada, Spain
gCardiff Catalysis Institute, School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff, CF10 3AT, Wales, UK
hUK Catalysis Hub, Research Complex at Harwell, Rutherford Appleton Laboratory, R92, Harwell, Oxfordshire OX11 0FA, UK
iUniversità degli Studi di Milano, Dipartimento di Chimica, Via Golgi, 19, 20133 Milano, Italy
First published on 25th June 2025
We report the rapid microwave-assisted solvothermal synthesis of a Cu-MOF (metal–organic framework) with open metal sites, focusing on understanding its CO2 capture properties in relation to phase purity and stability. A combined experimental and theoretical approach is used to identify the MOF's structural features involved in the adsorption process. Specifically, Cu(I) defects are found to play an important role in the CO2 adsorption process, with the Cu-1 sample, synthesized using an optimized ligand/Cu precursor ratio for the highest phase purity, exhibiting more abundant Cu(I) defects as well as the highest adsorption capacity. Grand canonical Monte Carlo simulations show that the Cu(I) sites exhibit a greater affinity for CO2 adsorption compared to the Cu(II) sites. In situ soft and hard X-ray absorption fine structure spectroscopic techniques confirm the conversion of Cu(I) to Cu(II) upon CO2 chemisorption, with this conversion being more pronounced in the core of the particles. The simulations are used to estimate the fraction of Cu(I) defects and Cu(II) sites present within the Cu-1 and to validate the experimental isotherm. Overall, this study provides insights into the CO2 capture properties of this type of Cu-MOF and highlights the importance of phase purity and the role of defects in achieving high adsorption performance.
In the field of CO2 adsorption, an archetypal example is HKUST-1 (also denoted as Cu3(BTC)2, BTC = benzene-1,3,5-tricarboxylate).11 In HKUST-1, OMS can be easily generated by removing the apical solvent molecules with gentle heating under low pressure, resulting in the activated structure with exposed Cu2+. It is proposed that CO2 can interact with the Cu2+ OMS in HKUST-1 after activation and removal of solvent molecules.12 Moreover, recent evidence also suggests the involvement of defective Cu+ sites in the reversible adsorption of CO2.13
Amination is another strategy to introduce NH2 functionalities in a solid adsorbent, which are selective for CO2 capture.14 By using 5-aminoisophthalic acid as the ligand under solvothermal conditions of a mixture of water, ethanol and dimethyl formamide (DMF), Zhao et al. synthesised a Cu-based MOF, namely JUC-141, which was found to have high CO2/N2 selectivity values as well as a high adsorption enthalpy when compared to other MOFs.15 This MOF is constructed using 5-aminoisophthalic acid (H2AIPA) coordinated to Cu2+ cations via the carboxylic groups to form the classic paddle wheel secondary building units (SBU) and with the amino groups linked to the dipole of the paddle wheel to form a 3D porous framework.15 The higher adsorption enthalpy is attributed to the strong interactions of CO2 with the amino groups due to confinement effects within the channels.15
A similar Cu(II) MOF based on the 5-aminoisophtalic acid ligand (Cu(AIPA)) was synthesised by Xu et al.16 via solvothermal synthesis using only DMF as the reaction solvent. This structure contains Cu paddle wheel units, similar to HKUST-1 and JUC-141, but now the amino groups belonging to the organic linkers are free and protrude into the pores of the system (Fig. 1), leading to a monoclinic structure. After removal of the DMF molecule coordinated to the Cu-paddle wheels, this synthesis afforded a MOF structure that is thermally stable above 500 K for over 36 h, with an interesting adsorption ability towards formaldehyde.16
In this work, we focus on the monoclinic Cu(AIPA) as a CO2 adsorbent because the bimodal chemisorption sites present allow, in principle, selective gas sorption. The CO2 adsorption capacity of this MOF was reported to be very similar to the value for JUC-141 (70 cm3 g−1 and 79.95 cm3 g−1 for Cu(AIPA)16 and JUC-141,15 respectively) but the surface area of the latter is approximately one order of magnitude higher, whereas the pore size is larger (14 Å in diameter for JUC-141 and 9.9 Å × 9.2 Å for Cu(AIPA)), which would be consistent with the hypothesis that the metal dimers play a primary role in the adsorption properties of this MOF towards CO2.
The phase composition obtained is crucial for the adsorption properties of MOFs; however the impact of impurities on these properties has not been thoroughly investigated. Previous studies15,16 demonstrate that the synthesis conditions have a profound influence on the product. For instance, variations in these conditions can lead to the formation of JUC-141, where the amino groups of the 5-aminoisophtalic acid ligands are involved in coordination, or GIF-KUC, where these groups remain uncoordinated. Ahnfeldt et al.17 reported a systematic analysis for an Al-MOF, examining how factors such as the molar ratios of the metal salt to ligand, the choice of solvents, and the type of aluminum salts affect the resulting structure.
Their study revealed a strong sensitivity of the synthesis process to reaction conditions, leading to the formation of multiple known and unidentified phases across different compositions. This inherent complexity in phase behavior is likely to become even more pronounced during scale-up, potentially compromising the consistency of gas separation performance on a per-gram basis.
In this work, we report the optimization of a microwave-assisted solvothermal synthesis route for Cu(AIPA) metal–organic frameworks (MOFs), supported by a comprehensive structural analysis using both soft and hard X-ray absorption spectroscopy techniques, alongside powder X-ray diffraction. This approach enables the rapid and reproducible synthesis of phase-pure Cu(AIPA) MOFs under relatively mild conditions, while also leading to the formation of a wide variety of defect sites within the framework. Furthermore, we investigate the nature of the CO2–MOF interaction through in situ X-ray absorption spectroscopy techniques, providing direct insight into the local coordination environment of the active sites. Complementary molecular modeling is employed to elucidate the underlying mechanisms of CO2 capture at the atomic level, offering a detailed understanding of adsorption phenomena that are otherwise inaccessible through experimental methods alone.
Sample notation | Synthesis conditions | |||||||
---|---|---|---|---|---|---|---|---|
Cu![]() ![]() |
Cu(NO3)2·3H2O (mg) | 5-AIPA (mg) | DMF (mL) | T (K) | Reaction time (min) | Heating time (min) | Yielda (mg) | |
a Using Cu2(AIPA)2(DMF)2 as the empirical formula of an ideal Cu(AIPA) MOF (molecular weight: 635.58 g mol−1), and considering AIPA (250 mg) as the limiting reagent, the theoretical yield is calculated to be 0.439 g. The actual yield of the nearly pure Cu-1 phase corresponds to 19.8%. If other phases are present, the so-calculated percentage yield is overestimated. | ||||||||
Cu-1 | 1.6 | 400 | 250 | 10 | 373 | 30 | 10 | 87 |
Cu-2 | 2.2 | 550 | 250 | 10 | 373 | 30 | 10 | 134 |
Cu-3 | 2.8 | 700 | 250 | 10 | 373 | 30 | 10 | 170 |
Cu-B1 | 1.6 | 400 | 250 | 10 | 373 | 30 | 10 | n.d. |
Cu-B2 | 1.6 | 400 | 250 | 10 | 373 | 30 | 2 | 84.8 |
Cu-B3 | 1.6 | 400 | 250 | 10 | 373 | 30 | 5 | n.d. |
Cu-B4 | 1.6 | 400 | 250 | 10 | 373 | 30 | 20 | 91 |
Cu-B5 | 1.6 | 400 | 250 | 10 | 373 | 30 | 1 | n.d. |
Cu-B6 | 1.6 | 400 | 250 | 10 | 373 | 30 | 10 | 81 |
The reproducibility of the synthesis was thoroughly evaluated, and consistent trends in phase purity were confirmed as a function of the Cu:
AIPA molar ratio. Additionally, we investigated the influence of microwave heating time and specifically, the time to reach the target temperature. These parameters were found to significantly affect the phase composition and defect distribution, providing further insight into the tunability of the synthesis process. For Cu-1, Cu-2 and Cu-3, the reactor was ramped up to the targeted temperature in 10 minutes. B1, B2, B3, B4, B5 and B6 represent a replica of Cu-1 but the reaction temperature was reached in 10, 2, 5, 20, 1 and 10 minutes, respectively.
To further remove DMF, an additional step of washing was performed on a set of samples. Accordingly, the solids recovered by filtration after the hydrothermal synthesis were washed with 10 mL of methanol three times and then dried under vacuum at 343 K for 12 hours. These samples are referred to as Cu-1meth, Cu-2meth, and Cu-3meth. A Cu-BTC, otherwise named HKUST-1, was prepared for comparison as follows: 3.38 g of Cu(NO3)2·3H2O (Sigma Aldrich 98%) was dissolved in 75 mL of distilled water and stirred until a clear solution was obtained. 2.94 g of benzene-1,3,5-tricarboxylic acid (BTC) (95% purity, Sigma Aldrich) was dissolved in 75 mL of ethanol (Sigma Aldrich, 99.5%) and stirred until a clear solution was obtained. The two solutions were mixed, and the resulting mixture was thermally treated at 403 K for 30 minutes using the microwave synthesizer (Discover SPCEM), operating at 200 W in a hermetically sealed glass vial. The solid recovered by filtration was washed with 10 mL of distilled water (three times) and then dried under vacuum at 373 K for 12 hours.
The isosteric heat of adsorption (Q) was calculated using DL-Monte as follows:
![]() | (1) |
To evaluate gas separation performance, we calculated the selectivity S(i/j) of Cu-MOF for binary and ternary gas mixtures using:
![]() | (2) |
Before the dynamic CO2 breakthrough measurement, approximately 400 mg of the sample powder was placed in a stainless steel tube (i.d. 5.0 mm, length 50 mm) and activated at two different temperatures (413 K and 493 K) with helium flow (20 mL min−1) for 1 hour. Once the sample was activated, the designated gas mixture N2:
CO2, 85
:
15 (20 mL min−1) with different relative humidities (0/85%) was monitored until equilibrium was reached. In addition, the sample column was conditioned at two different temperatures (273 K and 298 K) to study its influence on the CO2 capture properties of the material. Finally, the gas flow was switched to the column reactor for the measurement that was considered to be completed when the detected CO2 concentration reached the initial values.
The isosteric heat of adsorption (Q) is determined from the experimental CO2 adsorption isotherms measured at different temperatures using the Clausius–Clapeyron equation. Initially, adsorption isotherms are experimentally measured at two temperatures, 273 K and 298 K, using a Micromeritics 3Flex volumetric instrument, where the amount of CO2 adsorbed is recorded across a wide pressure range. Specific adsorption amounts are then selected for analysis, typically 0.10, 0.25, 0.50, 0.75, and 0.9mmol. For each temperature, the pressure corresponding to the chosen adsorption amounts is identified, resulting in pairs of (P,T) values.
Subsequently, the Clausius–Clapeyron equation is applied by plotting the natural logarithm of pressure (lnP) against the reciprocal temperature (1/T) for each selected adsorption amount. The resulting ln
P vs. 1/T plot yields a straight line, where the slope is proportional to the isosteric heat of adsorption (−Q/R), with R being the universal gas constant. Finally, the slope of the linear plot for each adsorption amount is used to estimate Q, providing insights into the temperature dependence of the adsorption isotherms. The CO2 sorption isotherm was also carried out for selected samples at 273 K on the Quadsorb Instruments analyzer located at the Research Complex at Harwell, with the pressure ranging from 0 to 100 kPa. Prior to the analysis, the samples (approximately 100 mg) were degassed for 2 hours at 473 K and 1 Pa. Helium was used for free space corrections and an isothermal bath was used to adjust the sample temperature during measurements.
A linear background subtraction was carried out before intensity normalization of the spectra in the energy window above the absorption resonances (965–980 eV). To study the beam damage of the samples, Cu L-edges NEXAFS were measured sequentially on the same spot using the fast and slow-motion mode of the monochromator (Fig. S2, ESI†). We note a non-negligible effect of the beam with time, especially in the slow mode. However, the first measurements recorded in the fast scanning mode of the monochromator can be safely considered representative of the actual initial structure and therefore these are used herein to discuss the impact of synthesis conditions on the electronic structure characteristics of the samples.
Soft XAS measurements under a CO2 atmosphere were performed at the ISISS beamline at Helmholtz-Zentrum Berlin (HZB) using both TEY and Auger electron yield (EY). EY XAS spectra were recorded with an analyzer setting of 50 eV pass energy (Ep) and electron kinetic energies (KE) of 385 eV, 385 eV, 385 eV, and 240 eV for Cu L-, O K-, N K- and C K-edges, respectively. The beam-line settings were exit slit (ES) 111 μm × 180 μm and cff 2.25. XAS spectra in total electron yield (TEY) were recorded using a Faraday cup positioned near the sample in the APXPS chamber.
Furthermore, spatially resolved X-ray absorption near edge spectroscopy (XANES) and X-ray fluorescence (XRF) measurements were performed at the I14 hard X-ray nanoprobe beamline at the diamond light source.37 For operando experiments, the sample was placed in a gas environment with heating element (DENS solutions),38 which was positioned on a kinematic mount. N2 was flown through the cell at a rate of 0.449 mL min−1 and CO2 was flown through at a rate of 1 mL min−1. Point XANES spectra at the I14 beamline were obtained ex situ, by placing the sample on a silicon nitride window at room temperature and exposing to air. XRF data were collected using a four-element silicon-drift detector in the backscatter geometry (RaySpec, UK). By scanning the X-ray energy and collecting XRF maps at multiple energies through the Cu absorption edge XANES maps were extracted. A dwell time of 10 ms per pixel per energy step was used for both the XRF and XANES maps, and dwell time was varied between 10 ms and 100 ms per energy step for the point XANES scans. XRF imaging was performed at 9.5 keV for measurements corresponding to point XANES measurements, and at 9 keV for measurements corresponding to the XANES maps. The size of the beam at the sample position, and therefore the spatial resolution of the scans, was 50 nm in diameter. Principal component and cluster analysis of the XANES scans was performed using MANTiS software.39
![]() | ||
Fig. 2 Rietveld multiphase refined PXRD and SEM in the back scattering electron mode of Cu-1 (a and b); Cu-2 (c and d); and Cu-3 (e and f). Inset in (f) is a SEM in secondary electron mode. |
The resonance B is not found in the bulk of Cu-2 and is abundant in Cu-3; it is present in Cu-1 as well. A more intense B resonance is also formed in Cu-2meth due to the methanol washing and heating treatment. The disappearance of the feature B upon thermal activation of HKUST-1 was associated with a thermally induced change of the Cu(II) coordination geometry, probably by the desorption of the solvent molecules.18 The resonance D at 938.2 eV was found for H-adsorbed Cu(I) species,42 but it is also due to the 2p → 4s electronic transition in the Cu2+ ions. Most of the changes observed upon washing involve the bulk of the samples with a reduction of the abundance of Cu(I) sites in Cu-1meth, whereas these are formed in the case of the Cu-2meth, consistent with the Cu K-edge XAFS data. Changes in the abundance of Cu(I) sites upon washing are intuitively linked to interaction of the solvent molecules (DMF or methanol) with these uncoordinated defective sites. In the case of the Cu-3meth, no changes in the abundance of Cu(I) sites are seen, but the B resonance is now less visible, with a spectrum very similar to Cu-2. It is interesting to note that structural similarities between these two samples (Cu-2 and Cu-3meth) were observed also in the EXAFS data (between 1.8 and 2.6 Å in Fig. S5c, ESI†) and related to secondary phases.
The analysis of the N K edge (Fig. 3e) reveals not only the π* resonances at 401.2 and 403.3 eV of DMF43 but also a strong resonance at 405 eV attributed to NOx species44 more pronounced for the Cu-2 and Cu-3 samples. This species is related to the Cu2(OH)3NO3 impurity phase more abundant on these samples. An additional π* resonance at 399 eV in all the samples is due to –NH2 groups of the AIPA ligand. After washing with methanol and thermal treatment, the resonance at 405 eV in Cu-2meth and Cu-3meth is not as intense as in Cu-2 and Cu-3, respectively, indicating a successful removal of this NOx-containing phase. The Cu-2meth spectrum is very similar to those of Cu-1meth and Cu-1, consistent with the other structural characterization data presented above. We note however that not all DMF was removed by the washing procedure. This indicates that DMF is actually strongly bound to some of the Cu sites and not fully removed during washing with methanol and vacuum desorption at 403 K. The O K-edge spectra of these samples (Fig. S5a, ESI†) are very similar to the spectrum reported for HKUST-1.45 Two regions are identified: the sharp π* region around 532 eV due to 1s → O2p–Cu3d transitions and the broad σ* region around 540 eV due to 1s → O2p–Cu4sp transitions. The resonances in the intermediate region between 534 and 540 eV are due to C–O species. A similar conclusion can be drawn from the analysis of the O K-edge spectra of Cu-1 and Cu-2, with the latter one showing a much more intense π* resonance at 532 eV attributed to the Cu–O bond, possibly due to a bond N–CO–Cu of coordinated—DMF42 or due to the impurity Cu2(OH)3NO3 phase. The O K-edge NEXAFS spectrum of Cu-2meth resembles the spectrum of Cu-1, as expected when the abundance of the inorganic impurity phase is reduced.
In the C K-edge spectra in Fig. 3f, the expected contributions are from the ligand AIPA and the Cu-coordinated DMF for Cu-1, Cu-2 and Cu-3. For comparison we report here also the CK-edge of a Cu-BTC MOF, which does not contain DMF, as it was synthesised in an ethanol/water mixture. The main resonances found in Cu-BTC are assigned to transitions into the unoccupied π* orbitals of the phenyl system (∼285 eV) and the carboxylate/carboxylic acid groups (∼288 eV), similar to literature values for BPTCA.46 The peak at 288 eV is also the most intense one in DMF and attributed to transition into the –N–CO π* orbital. We can see that the relative intensity of these peaks is very similar for Cu-2, Cu-2meth and Cu-1meth, whereas Cu-1 presents a more intense π* orbital of the phenyl group amongst the samples. Since the washing with methanol removes this component at 288 eV for Cu-1/Cu-1meth, whereas the structure is maintained, we postulate that the surface of the Cu-1 is covered by carbonaceous species due to the residual, decarboxylated, uncoordinated ligand. This would be expected considering that the relative amount of AIPA used during this synthesis was the highest amongst the synthesis compositions investigated. Furthermore, we would like to understand if the removal of DMF can lead to structural distortion as indicated by the B resonance in the Cu L3-edge spectra. An interesting aspect is evidenced with the analysis of Cu-2 and Cu-2meth. When we compare Cu-2 with Cu-2meth in Fig. 3f, the ratio between the two resonances (phenyl system (∼285 eV) and the carboxylate/carboxylic acid groups (∼288 eV)) remains almost unaltered after washing; therefore, we could assume that a negligible amount of DMF was removed from the surface/near surface region by this treatment and the signal of the C K-edge in this region is dominated by the ligand contribution. The interesting finding is related to the transitions around 288.3–288.7 eV, consistent with transitions of the type 1s → π* of C-NH2 orbitals in the ligand, which are present with a relatively higher intensity in Cu-1 and Cu-1meth than in Cu-2 and Cu-2meth, whereas they are totally absent in Cu-BTC. We compare the Cu-2meth and Cu-1meth because these have a similar bulk structure (FT-EXAFS in Fig. 3c and XANES in Fig. S5c and d, ESI†) and similar content in terms of the nitrate impurity phase (NK edge spectrum in Fig. 3e). These two samples have the same intensity of ligand resonances (phenyl and carboxylates) but the 1s → π* C-NH2 resonance is lower for Cu-2meth. This can be attributed to either its decomposition during methanol washing/heating or a loss of direct or solvent-mediated coordination to Cu2+. According to the Bly-holder model,47 the resonance intensity in absorption spectroscopy is related to electronic effects induced by the ligation of the ligands L to the metal center M, with increase of the intensity of the resonance for a L to M π-donation, whereas a decrease in intensity occurs when also a π-back donation from M to L takes place.47 It follows that such a coordination of the C–NH2 moieties would be even more pronounced for the Cu-2, consistent with the fact that Cu-2 has a similar first and second coordination shell to the Cu-BTC (Fig. 3b). Such a site is however unstable during the thermal treatment of the washing process, which removes the coordination to NH2 leaving Cu(I) sites in the main monoclinic Cu(AIPA) phase. Based on the similar spectrum of the Cu-2 with the Cu-BTC one might speculate that the NH2 moiety of the AIPA ligands might interact with the Cu sites of the paddle wheels, generating a saturated Cu(II) coordination environment similar to HKUST-1 (Fig. S6c, ESI†). The morphological characteristics of this sample would suggest the presence of another phase, which however we were unable to assign (Fig. 2c and d). The thermal treatment following the methanol washing procedure is effective for the removal of this phase leading to a phase transformation that stabilizes the Cu(AIPA) (CCDC, deposition number 628816) phase, in which the AIPA-NH2 group is not interacting anymore with Cu sites (Fig. S1c, ESI†), leading however to the formation of Cu(I) defective sites (resonance C in Fig. 3d) and other structural distortions (resonance B in Fig. 3d). This additional coordination could also explain the additional peaks in the diffractogram of Cu-2, but we have not pursued this further. Cu-3 exhibits greater structural complexity and heterogeneity. The lower intensity of the C K edge resonances for the Cu-3 is consistent with a surface dominated by the inorganic NOx-containing phase as expected from the highest amount of Cu nitrate precursor used during the synthesis of this sample. Based on these results, we postulate that, during the several synthesis steps, the favored coordination of Cu(II) to carboxylate species of the ligands ultimately drives the phase transformations towards the target phase. However, when Cu(II) species are in excess –NH2 groups remain coordinated hindering the crystallization of the target phase.
Sample | Cu-1 | Cu-1meth | Cu-2 | Cu-2meth | Cu-3 | Cu-3meth |
---|---|---|---|---|---|---|
a Samples were subjected to degassing at 413 K for 3 h at 1 Pa before N2 physisorption. | ||||||
SBET/m2 g−1 | 267 | 349 | 191 | 83 | 286 | 71 |
Pore volume (DFT method)/cm3 g−1 | 0.238 | 0.32 | 0.203 | 0.085 | 0.215 | 0.079 |
Pore size diameter (DFT method)/nm | 5.88 | 4.15 | 4.88 | 5.68 | 4.89 | 2.77 |
The monoclinic Cu(AIPA) structure was described as formed by 2D sheets generating a 3D framework through non-classical hydrogen bond interactions between C–H groups in a benzene ring from one layer and a doubly bridging carboxyl oxygen atoms from the adjacent layer. Thus, parallel layers are coupled through these weak hydrogen bond interactions, forming a 3D supramolecular framework. We suggest that the local disorder of the H bond networks between layers is responsible for the larger pore size obtained and thus a lower surface area. The shape of the isotherm changes sharply with the degassing pre-treatment (Fig. S7b, ESI†) and provides more insights into the porous characteristics of these samples. At the lower degassing temperature (393 K, overnight) corresponding to very low surface areas as summarized in Table S6 (ESI†), the isotherm presents the type H3 hysteresis typical of slit-shaped meso- and macropores in materials with a plate-like layered structure, whereas the micropores are not accessible during our experiments. When the degassing is performed at higher temperature (413 K, 3 hours), the isotherm presents multistep-like behaviour in both the adsorption and desorption branches and a larger hysteresis loop, which is generally observed in highly flexible aggregates of plate-like particles48 or when the macropores are filled with pore condensates. It was also reported for pore structures containing both open and partially blocked mesopores.48 The abrupt changes observed in the desorption branch of the isotherms of these samples indicate a layer-to-layer structural instability during the adsorption experiments, consistent with higher interlayer flexibility due to a more disordered H-bonding network between the layers. The resulting surface area and in Table 2 is very similar for Cu-1 and Cu-3 and slightly lower for Cu-2. We should also note that only Cu-1 is an almost pure monoclinic Cu(AIPA) phase with larger primary particles free from deposits of secondary phases, whereas Cu-2 and Cu-3 present additional phases deposited on the primary particles, which will contribute to the adsorption behaviour.
Washing with methanol followed by thermal treatment was effective to a certain extent to remove carbonaceous impurities as well as accessible DMF molecules as shown with an increase of the surface area and pore volume for the Cu-1, Cu-1meth system, confirming that this is the most structurally stable sample. In contrast, the BET surface area and pore volume are significantly reduced upon methanol treatment for Cu-2/Cu-2meth and Cu-3/Cu-3meth systems, indicating the poor stability of the other porous unknown phases contained in these samples and the more defective Cu(AIPA) phase. The thermogravimetric (TG) analysis of Cu-1 in Fig. S8a (ESI†) shows the DMF removed by methanol washing, amounting to 2% by weight. The TG analysis (Fig. S8b, ESI†) for Cu-2 indicates the presence of water in the pores as well as DMF, accounting for the 19% by weight. The washing step with methanol enables a more effective exchange of the DMF, leading to a very similar residual DMF content of this sample to Cu-1meth, around 10%. However, this results in a loss of more than half the surface area, which means that these impurity phases, different from the monoclinic Cu(AIPA), are not suitable for gas adsorption studies. Interestingly, on Cu-3meth, the amount of weakly chemisorbed adsorbates increases, indicating a further phase transformation generating adsorbed water. The TG analysis of these samples indicates that DMF desorption occurs in the range between 463 and 513 K, whereas the onset temperature at which the structure collapses is 533 K followed by two thermal processes with maximum rates at 547 K, 557 K and 563 K (Fig. S8, ESI†). Samples containing impurities such as Cu-2 and Cu-3 show a more complex thermal desorption profile, in which the first peak is less intense (Fig. S8, ESI†). This peak can be attributed to the decomposition of the carboxylate species of the ligands, which were used at higher concentrations in the synthesis, leading to the formation of the purest Cu(AIPA) phase.
The CO2 adsorption isotherms of Cu-1 are plotted in Fig. 4a. From the conditions we explored in this work, it is clear that the uptake capacity of CO2 for this sample depends on the activation pre-treatment as well as the temperature at which the adsorption is carried out. When the activation is carried out at 413 K and 0.1 Pa for 1 hour, the CO2 uptake at saturation is ∼1.86 mmol g−1 and 0.93 mmol g−1 at 273 K and 298 K, respectively, both measured at 100 kPa. When the activation is carried out at 493 K, the CO2 uptake at saturation is ∼1.46 mmol g−1 and 0.15 mmol g−1 at 273 K and 298 K, respectively, both measured at 100 kPa. The isosteric heat of adsorption presents a maximum value of 40 kJ mol−1, which is accounted globally as a physisorption process, and decreases slowly with the amount of CO2 adsorbed (Fig. S9, ESI†), indicating an energetically heterogeneous surface.49
The maximum adsorption uptake at 100 kPa is comparable to other MOFs previously reported under relatively similar conditions. For example, Zarate et al. have studied the adsorption properties of MIL-53(Al)-NH2 under 1 bar at 283 K and measured a capacity of 1.6 mmol g−1.50 Shekhah et al. reported a capacity of 2.3 mmol g−1 under 1 bar at 308 K for SIFSIX-3-Zn.51 Grajciar et al.52 measured a capacity at a saturation of 13 mmol g−1 under 15 bar at 303 K for Cu-BTC. Zhao et al. reported CO2 adsorption capacities for JUC-141 one order of magnitude higher compared to this study.15 It should be noted that a lower CO2 uptake should be considered as the result of a more disordered structure, which is expected for samples obtained via rapid microwave assisted solvothermal synthesis. A lower adsorption capacity in our studies when the activation is carried out at 493 K can be attributed to the highly flexible layered structure of this MOF and the easy generation of slit-shaped mesopores and larger macropores upon thermally-induced ligand detachment from the metal centre and the introduction of Cu(I) defects in the structure. We also observe that the degassing pressure used in the activation process has a strong influence on the final adsorption capacity where a higher pressure of 1 Pa, amongst other factors, yielded a much lower CO2 uptake of 0.4 mmol g−1 (Fig. S10a, ESI†).
This is consistent with TG, XRD (Fig. S11, ESI†) and N2 adsorption isotherm analysis after CO2 adsorption experiments (Fig. S12a and b, ESI†) and after degassing at 473 K and 1 Pa for longer time 12 hours (Fig. S12c, ESI†), showing that the structure is globally preserved although the specific surface area decreases due to a change in the 3D network of H-bonds induced by the thermal treatment and the evacuation process during the activation leading to an increase of the macroporosity. Nevertheless, in terms of selectivity, the GCMC simulations, presented later on, predict high enthalpies in the presence of defective sites, which makes this material a potentially interesting candidate for the separation of gas mixtures. This is hypothesised based on the generation of open Cu sites as well as Cu(I) defective sites. Consequently, we study the capture properties of these materials for CO2 in a simulated CO2/N2 gas mixture at ambient pressure. The breakthrough curves at 273 K and 298 K are reported in Fig. 4b and c, respectively. The amount of CO2 adsorbed from a 15:
85, CO2
:
N2 mixture at 273 K amounts to 0.62 mmol g−1 when the activation of Cu-1 is carried out at 493 K and to 0.41 mmol g−1 for activation at 413 K, indicating the beneficial effect of the higher degassing temperature on the formation of selective sites for CO2 capture, despite the overall specific surface area being decreased. The adsorption capacity for a simulated mixture is significantly reduced when the CO2 is adsorbed at room temperature. At 298 K, the amount of CO2 adsorbed from the mixture significantly decreases to 0.17 mmol g−1 for degassing at 493 K and to 0.09 mmol g−1 for degassing at 413 K. One could postulate that the remaining sites are the portion of highly selective sites for CO2 capture.
The presence of water molecules (relative humidity 85%) is however detrimental, leading to a significant decrease of CO2 adsorbed amounting to 0.4 mmol g−1 and 0.05 mmol g−1 at 273 K and 298 K, respectively. Consequently, despite the sufficiently good affinity for CO2 capture, the presence of water limits the applicability of these systems for selective capture.
![]() | ||
Fig. 5 NEXAFS spectra measured for Cu-1 under CO2 pressure in the millibar range as indicated: (a) Cu L-edges (order of the experiments from bottom to top) and (b) O K edge. |
The soft X-ray in situ study on Cu-2 did not show any dynamic changes at the Cu L-edges in both TEY and EY (not shown), the first slightly more bulk sensitive than partial Auger electron yield (EY), indicating the detrimental effect of the impurity phases (Fig. 2d). An explanation for this finding is that: (i) the nanostructures deposited on the surface covering the large particle are not involved in the CO2 adsorption and are detrimental by preventing CO2 access into the pores and (ii) the low abundance of Cu(I) sites for this sample. We performed an additional operando study on the Cu-1 at the I14 beamline at DLS by means of hard X-ray spectromicroscopy at the Cu K-edge with the goal to have information about the localization of the CO2 adsorption process. These data are reported in the ESI† in Fig. S14. Here, XANES maps are obtained under in situ conditions: first the sample is measured at room temperature while flowing N2, then at 473 K in N2 and lastly at room temperature while flowing CO2. Each of these maps was taken in a different area on the sample, but with some overlap between all three maps (Fig. S14a, ESI†). Principal component analysis was then used to identify clusters of XANES spectra, which are similar in shape and thus have a similar chemical composition (Fig. S14b, ESI†).
Accordingly, we can observe that the sample is composed of a mixture of Cu(I)/Cu(II) in N2 at room temperature (Fig. S14c, ESI†). To study differences in the ratio of Cu(I) to Cu(II) upon changes in the in situ conditions, we study the absorption intensity at the pre-edge peak with respect to the edge (Ipe = intensity pre-edge/intensity edge × 100) rather than performing a fit of reference compounds. This is the preferred method due to a lack of suitable reference samples that would be required for fitting the full spectra. For the initial sample conditions Ipe= 38% for the bulk of the sample, and Ipe = 40% for the edge.
Next a thermal treatment up to 473 K in N2 flow was performed to remove the DMF as found by TG analysis (Fig. S11a, ESI†). Under this condition, we can see no changes in the particle shape, whereas the Ipe increases to 75% of the edge intensity and the pre-edge peak moves to slightly higher energy (8980.6 eV to 8982.1 eV), indicating a change in copper speciation and/or the coordination environment of the Cu atoms. This process seems slightly more pronounced in the bulk than in the edge of the sample, where Ipe only rises to 71%. The increase in Ipe indicates an increase of Cu(I) abundance, corresponding to the formation of uncoordinated defective sites, but the energy shift would contradict this and instead indicate an increase in Cu(II). Therefore, we conclude that in addition to a speciation change the coordination environment of the Cu atoms is changed, affecting the spectral shape. This could be caused by the removal of DMF, as organic solvents can change the shape of the XANES spectrum53 or potentially by changes in the porosity of the material or the detachment of one of the ligands during thermal annealing. Possible effects from thermal degradation are further discussed below. Notably, by cooling down to RT and feeding CO2, we observe a decrease in the pre-edge peak (Ipe = 67% for the bulk, Ipe = 66% for the edge), which indicates a reduction of the Cu(I) resonance, consistent with the soft X-ray data. Whilst we were able to capture these dynamics, some thermal degradation of the sample at 473 K could be expected together with the possibility that the fluid dynamic conditions realized in the in situ cell limit the diffusion of the CO2 into the framework. Furthermore, effects of beam radiation are observed in the XRF image (Fig. S14a, ESI†); thus, a study of the influence of the X-ray beam on the XANES spectra is required.
To investigate the effect of beam damage on the hard X-ray XANES spectra, we obtained point-XANES measurements (Fig. S15, ESI†). First, a scan was taken with 10 milliseconds per energy step dwell time (scan 0), corresponding to the dwell time used to obtain the XANES maps in Fig. S14 (ESI†). Next, another scan was taken in a different position at 100 milliseconds per energy step dwell time (scan 1), corresponding to a 10× higher beam exposure than was used to obtain the XANES maps. A total of 4 consecutive scans were taken in this position, to investigate changes in the spectral shape upon repeated beam exposure. The mapped XANES spectra discussed before were the result of an average from each region, where each individual pixel contains one spectrum. In contrast, for the point XANES performed here, with a focused (50 nm) beam (Fig. S15, ESI†), this averaging cannot be performed, resulting in significantly increased noise levels in the spectra in the post-edge region (Fig. S15b, ESI†). This complicates normalization of the spectra, making a direct comparison of the full spectral shape of the various scans difficult. As such, we instead use the ratio between the intensity of the pre-edge peak and the absorption edge prior to normalization (Ipe). The results give Ipe = 37% for scan 0 and Ipe = 35% for scan 1. This indicates that during a single XANES scan beam damage does not significantly alter the shape of the absorption edge region regardless of dwell time, even if beam effects can be observed after the full scan is completed in corresponding XRF images (Fig. S15a, ESI†). This can be understood in that the scans are performed from low to high energy levels and beam effects will be strongest at energies that are equal to or higher than the absorption edge. In concurrence with this, when consecutive scans were taken in the same position as scan 1, Ipe rises to 47% or 48%, indicating that beam damage does play a role when repeatedly measuring the same area. As such, this raises questions on the reliability of the 473 K N2 and RT CO2 scans. However, it should be considered that point XANES measures continuously in one area, while XANES maps scan in a raster, allowing time for beam effects (such as local sample heating) to dissipate between scans. Additionally, consecutive scans here are performed at 10× the dwell time used in the XANES maps. The maps are also taken in different areas, with some overlap between all. This did not give additional clusters for the measured and unmeasured area, so we conclude that the spectra are unaffected by the beam.
In addition to beam damage effects, Cu MOFs can degrade when exposed to high temperatures for prolonged periods of time. This can further affect the shape of the XANES scan.54 The single point CO2 isotherms in Fig. 4 show that for the degassing at 0.1 Pa, the amount of CO2 adsorbed by Cu-1 is higher when the degassing is performed at 413 K compared to 493 K, which is a clear indication of a loss of exposed surface area and porosity due to a thermal induced partial structural degradation. As the XANES scan at 473 K takes 2 hours and 45 minutes to complete, this effect can be expected to occur during this measurement. Since the XANES scans were taken from low to high energy levels, the absorption edge of the 473 K measurement is reached after keeping the sample at this temperature for 1 hour, so that the comparison between edge and pre-edge peak intensity should remains valid. However, it can be expected that the spectral shape in the RT CO2 measurement is affected compared to the previous scans. This complicates any quantitative analysis of the spectra, but an increase in Cu(II) abundance is still apparent regardless of the formation of Cu(I) sites during thermal annealing.
We show here a reduction of Cu(II) to Cu(I) during heat treatment that is slightly more pronounced in the bulk of the sample compared to the edges. We also observe a subsequent oxidation back to Cu(II) upon CO2 capture. However, careful consideration is required with regard to thermal degradation of the sample, which can affect the spectral shape and therefore limits the analysis to qualitative rather than quantitative interpretation.
A slight discrepancy is observed between the theoretical and experimental isotherms (see Fig. 6), which could arise from several factors: (i) a minor deviation in defect concentration; (ii) limitations in the Cu-1 model, which assumes homogeneous defect distribution, whereas defect interactions may actually lead to a non-uniform distribution; (iii) inaccuracies in the force field parameters used; and (iv) limitations in the qEq approach for calculating the partial charges in the Cu-1 structure. These factors may contribute to minor inaccuracies and a slightly lower predicted adsorption capacity at saturation. Nonetheless, the good overall agreement between the experimentally measured and simulated isotherms provides sufficient confidence to infer key parameters that govern CO2 adsorption and its microscopic mechanism. We specifically examined the adsorption isotherm at low pressures and at saturation to gain insight into possible CO2 adsorption mechanisms at the microscopic level.
To examine the interactions at low pressure, we isolated and optimized three clusters containing the defective sites in each cluster (see Fig. S1d–f and S16a, c, e, ESI†) and used the adsorption locator approach based on the simulated annealing algorithm for locating CO2, H2O and N2 within each cluster.55,56 Analysis of these calculations shows clearly that the CO2 molecule is adsorbed in the order: Di-Cu(I) > Tri-Cu(I) > Tetra-Cu(II). The calculations indicated different adsorption energies for the different defective sites: −80 kJ mol−1, corresponding to the Di-Cu(I) site, −75 kJ mol−1, corresponding to the Tri-Cu(I) site and −31.5 kJ mol−1, corresponding to the Tetra-Cu(II) site, with the calculated Cu–O bond lengths of 1.90 Å, 2.15 Å and 2.75 Å, respectively (Fig. 7). The adsorption energy of CO2 at the Di-Cu(I) sites is more than twice that of Tri-Cu(II) and Tetra-Cu(II) sites. The CO2 adsorption enthalpy was investigated by Liang et al.57 for Cu-BTC (H-KUST-1), which contains the same copper paddle wheels. In their study, the enthalpy of adsorption determined directly by differential thermal analysis (DTA) was −30 kJ mol−1, consistent with our calculated value for the case of the Tetra-Cu(II) site. However, their work did not specify the nature of the sites where the CO2 molecules were adsorbed. Grajciar et al.52 studied CO2 adsorption in a Cu-BTC MOF by using DFT calculations and compared the results with microcalorimetry. They reported a heat of adsorption of −29 kJ mol−1 obtained experimentally and −29.1 kJ mol−1 calculated by the DFT method. In their theoretical model, they considered a very low loading of CO2 molecules, which are physisorbed on the Tetra-Cu(II) site. However, the values of the adsorption energies of a CO2 molecule in the cases of Di-Cu(I) and Tri-Cu(I) sites are 80 kJ mol−1 and 75 kJ mol−1, which can be considered as chemisorption. These relatively high heats of adsorption in the case of the Di-Cu(I) and Tri-Cu(I) defective sites can be attributed to a sterically favorable electrostatic interaction with CO2. Therefore we calculated the partial charges, which are 1.6 for q(Di-Cu(I)), 2.74 for q(Tri-Cu(I)), and 3.13 for q(Tetra-Cu(II)). The lower positive charge indicates that the copper contains higher electron density and might contribute to a stronger π-back donation from the orbitals of the metal ion to the orbitals of the CO2 molecule.58 Su et al.59 also demonstrated that the back-bonding interactions between metal orbitals and antibonding orbitals of small molecule guests promote the chemisorption process. Although the calculated adsorption energies do not explicitly account for π-back donation, which was shown to occur in Cu-(II)-paddle wheel MOFs,58 our results confirm that steric hindrance, site-specific interactions and lateral electrostatic effects play a significant role in influencing adsorption behaviour. These factors collectively affect the binding affinity and orientation of CO2 at the adsorption sites, underscoring the complexity of adsorption mechanisms. The predicted enthalpy of adsorption for CO2 interacting with a MOF containing copper metal, calculated using the GCMC approach, range from 29 to 75 kJ mol−1, while the experimentally measured enthalpy is calculated to be between 25 and 40 kJ mol−1.
This discrepancy between the theoretical predictions and experimental results can be reasonably explained by considering the limitations of the GCMC approach and thermodynamic factors. The GCMC approach is a powerful simulation technique widely used for modeling adsorption processes in porous materials. However, it has some inherent limitations that may affect the accuracy of the predicted adsorption enthalpies likely due to force field limitations, and the treatment of defects. The adsorption enthalpy reflects the overall energy change when CO2 molecules interact with the Cu-MOF surface. Several factors can influence the difference between the predicted and experimental values. Experimentally, the enthalpy of adsorption is often determined as an average over multiple adsorption sites, including both strong and weak binding locations. The experimentally measured range (25–40 kJ mol−1) probably reflects a combination of various adsorption sites, including those at defect-free and defect-containing regions (see Fig. S9, ESI†). The broader range predicted by GCMC (30–75 kJ mol−1) may indicate that the simulation captures stronger interactions at specific defect sites, while less accurately representing the weaker interactions. The actual concentration and distribution of defects in the Cu-MOF can vary, affecting the average adsorption strength. If the GCMC simulations assume a higher number of defect sites than that is present experimentally, then this may lead to overestimated enthalpies. In contrast, experimental samples that might have a more diverse range of local environments result in a lower observed average. Our calculated values are consistent with the observations of Drenchev et al.,60 who studied open metal sites in the organic framework CPO-27-Cu (Cu MOF 74), containing a paddlewheel unit, which is structurally similar to the Cu-MOF, and deduced from IR spectroscopy an enthalpy of adsorption of about 63 kJ mol−1 for a CO2 molecule specifically adsorbed on the Tri-Cu(I) site, and approximately 20 kJ mol−1 on the Tetra-Cu(II) site. Itadani et al.61 studied the coordinative unsaturated Cu(I) sites in the MFI zeolite and found high isosteric heats in the range of 50–130 kJ mol−1 for CO2 molecules. The most probable arrangements of CO2 were further investigated by GCMC simulations under saturation conditions and the main adsorption sites were extracted (see Fig. S17a–e, ESI†). It can be seen that the simulated isotherm reproduces well the experimental data, corroborating the validity of the microscopic models of the CO2 adsorption mechanism. An in-depth analysis of the snapshots at saturation allowed us to gain more insight into the configurations of the adsorbed molecules. In the case of the Tetra-Cu(II) site, the CO2 molecule is oriented with the O atoms pointing towards the Cu(II) ion of the framework at an average distance of 3.00 Å from the Tetra-Cu(II) and of 3.45 Å from the O atom of an adjacent CO2 molecule in closest proximity. In the case of the Tri-Cu(I) site, the CO2 molecule interacts via its O atoms with Cu(I) ion with a mean separating distance of 1.97 Å; the nearest CO2 molecule is located 3.54 Å away (intermolecular O–O distance). In the case of the Di-Cu(I) site, the CO2 molecule interacts via its O atoms with Cu(I) ion with a mean separating distance of 1.83 Å. That site has the potential to be the most favourable for CO2 adsorption. We also found a new adsorption site for CO2 molecules at saturation, which is the interaction between CO2 molecules and the –NH2 group (see Fig. S17c, ESI†), and the –NH3+ group (see Fig. S17e, ESI†) whereas, at low pressure, location at this site has not been observed. The calculated intermolecular O–H hydrogen bond length is 2.34 Å and 2.026 Å, between NH2–CO2 and NH3+–CO2, respectively, indicating that the interaction between the adsorbed CO2 molecule and the –NH3+ and –NH2 groups actively contributes to the process of adsorption of CO2. We suggest that the adsorption of the CO2 molecule on the –NH2 and –NH3+ groups can take place only when the coordinative unsaturated Cu sites are fully saturated by CO2 adsorption.
The GCMC simulations conducted in this study reveal that under-coordinated open metal sites create unique adsorption environments that increase the affinity and selectivity of the material for certain gas molecules. In the context of gas separation, tuning the concentration of such under-coordinated Cu(I) sites could by modulating the synthesis conditions provide a powerful tool for tailoring the selectivity of Cu-MOF materials towards CO2 capture from flue gas streams or natural gas purification. Defect engineering, such as deliberate introduction of structural vacancies or alterations in linker types, may further promote the formation of low-coordination metal sites, thus enhancing the tunable adsorption properties. The GCMC simulations were conducted on Cu-MOF structural models to evaluate the adsorption properties of CO2/N2 mixtures at a 15:
85 ratio. This specific gas mixture composition was chosen to mimic typical conditions encountered in gas separation processes, where the selective capture of CO2 from nitrogen-rich mixtures is critical for applications such as carbon capture and flue gas purification. The simulations were designed to assess the MOF's inherent selectivity for CO2 over N2 and to understand the role of specific structural features of the MOF in governing this behavior. To further investigate the impact of water on adsorption performance, the study extended the analysis to ternary mixtures by introducing H2O alongside CO2 and N2. The presence of H2O in the adsorption environment is highly relevant for real-world applications, as many industrial gas streams contain significant water content. Water molecules can compete with CO2 and N2 for adsorption sites, potentially altering the selectivity profile of the MOF and affecting the efficiency of gas separation processes. Therefore, a comparison was made between the binary mixture data for CO2/N2 (15
:
85) and the ternary mixture data for CO2/N2/H2O to elucidate the influence of water on the selective adsorption characteristics of the Cu-MOF (Table 3).
Cu-MOFs | Pressure (bar) | Temperature (K) | IAST selectivity of CO2/N2 (15![]() ![]() |
IAST selectivity of CO2/N2/H2O |
---|---|---|---|---|
Defective model 1 | 1 | 298 | 13 | 0.8 |
Defective model 2 | 1 | 298 | 9 | 0.5 |
The selectivity of the two structural models was evaluated at 298 K and 1 bar. The defective model 1, with a composition of 30% Tri-Cu(I), 65% tetra Cu(II), and 5% Di-Cu(I), exhibited a selectivity of 13. In contrast, defective model 2, consisting of 35% Tri-Cu(I), 55% Tetra-Cu(II), and 10% Di-Cu(I), showed a selectivity of 9. These results suggest that the selectivity is primarily governed by the chemical environment of the Cu sites in the Cu(AIPA), rather than by porosity, which appears to have an insignificant impact. As illustrated in Fig. 7, the CO2 molecules exhibit a stronger affinity for the Cu-MOF sites compared to N2, as evidenced by shorter intermolecular distances between CO2 and the Cu sites. However, the calculated enthalpy of adsorption for N2 molecules shows higher values at certain defect sites: −120 kJ mol−1 for the Di-Cu(I) site, −25 kJ mol−1 for the Tri-Cu(I) site and −160 kJ mol−1 for the Tetra-Cu(II) site. This apparent discrepancy arises because enthalpy values represent the energy released during adsorption but do not account for the spatial arrangement or the frequency of favorable interactions across the system. While N2 shows higher adsorption enthalpies at specific sites, these sites are less accessible compared to those favoring CO2 adsorption, resulting in lower overall selectivity for N2. CO2 has unique molecular properties, such as its linear structure and significant quadrupole moment, allowing it to engage in stronger and more directional interactions with unsaturated copper sites. These are dominated by electrostatic and Lewis acid–base interactions, leading to CO2 being localized closer to the copper centers. In contrast, N2, with its weaker quadrupole moment and less directional interactions, requires more energy to adsorb at certain high-energy sites but does not achieve the same proximity as CO2. The higher enthalpy values for N2 adsorption can also be attributed to the limitations of the GCMC approach, which, while valuable for modeling adsorption processes, is less suited for predicting accurate adsorption enthalpies. GCMC relies on empirical force fields that cannot fully capture electronic effects such as orbital overlap or charge transfer, which are essential for understanding the chemical nature of gas-MOF interactions. In contrast, density functional theory (DFT) provides site-specific adsorption energies and accounts for these effects, offering a more precise understanding of adsorption and energetic mechanisms. Therefore complementary DFT studies might enable to achieve a more accurate depiction of the adsorption and energetic processes at play.
Notably, the most favorable site for N2 adsorption is the Tri-Cu(I) site. However, the overall preferential adsorption of CO2 arises from its ability to interact more effectively with a broader range of sites, emphasizing the Cu-MOF inherent selectivity for CO2 over N2.
The ‘defect1’ model exhibits a notable abundance of Tri-Cu(I) sites, significantly enhancing its selectivity toward N2. When simulating N2/CO2 separation in the presence of water, both models showed a drastic reduction in selectivity, decreasing by a factor of 10. The results, summarized in Table 3, reveal clear differences in selectivity when H2O is introduced into the system. CO2/N2 selectivity decreases significantly in the presence of water, indicating that H2O molecules have a strong affinity for the adsorption sites within the Cu-MOF framework. This effect is probably due to the polar nature of water, which enables it to interact strongly with the metal centers and other polar sites within the MOF. These interactions can occur through strong hydrogen bonding with under-coordinated metal sites (see Fig. 7), which may preferentially bind H2O over CO2 or N2. Consequently, the CO2 adsorption capacity is reduced in the ternary mixture, as water occupies critical adsorption sites, thereby impacting the overall gas separation efficiency. The results highlight the need to consider the effects of water in practical applications, as the presence of water can reduce CO2 capture efficiency and selectivity.
The good agreement between experimental and computational data confirms that the presence of Cu(I) defects significantly improves CO2 adsorption capacity compared to defect-free analogues. Importantly, the ability to induce and tune these defects through rapid synthesis and post-synthetic thermal treatment offers a scalable and energy-efficient alternative to conventional solvothermal methods, with direct implications for industrial gas separation technologies. However, the presence of water remains a critical challenge, as it competes for adsorption sites and diminishes CO2 selectivity, besides causing structural degradation. This highlights the urgent need for the development of water-tolerant MOF systems. Future work should focus on engineering MOF-based composites and implementing strategies to mitigate water interference, thereby preserving high CO2 selectivity under humid conditions. These advancements will be essential for translating MOF-based materials into viable solutions for real-world carbon capture and gas purification applications.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ma00666j |
This journal is © The Royal Society of Chemistry 2025 |