Barbara Panić,
Tea Frey,
Mladen Borovina
,
Petra Ištoković,
Ivan Kodrin
* and
Ivana Biljan
*
Department of Chemistry, Faculty of Science, University of Zagreb, Horvatovac 102a, Zagreb, HR-10000, Croatia. E-mail: ikodrin@chem.pmf.hr; ibiljan@chem.pmf.hr
First published on 29th April 2025
We synthesized a series of azo-linked porphyrin-based porous organic polymers (APPs) with linear, bent, and trigonal linkers (APP-1 to APP-6) and with directly connected tetraphenylporphyrin units (APP-7a, APP-7b and APP-8). The synthesized APPs are amorphous solids demonstrating good thermal stability and diverse BET surface areas. APPs with linkers showed significantly higher surface areas (469 to 608 m2 g−1) compared to those with directly connected tetraphenylporphyrin units (0.3 to 23 m2 g−1). Higher surface areas correlated with enhanced CO2 adsorption, particularly for APP-1, APP-2, and APP-5 with experimental CO2 uptake values of 41 mg g−1, 38 mg g−1, and 38 mg g−1, respectively, at 306 K. The computational study supported the experimental findings and provided insights on how surface area and the local landscape affect the CO2 adsorption. Although the computational models were based on ideal structures, while the experiments revealed the materials were amorphous, the calculated CO2 adsorption capacities were roughly comparable to the experimental results, particularly for the 3D systems (APP-5 and APP-6) and the 2D systems with directly connected building units (APP-7 and APP-8). Porphyrin units in the framework serve as additional binding sites for CO2, especially when unhindered and available on either side of the porphyrin plane. This work highlights the potential of 2D layered APPs and 3D topologies for efficient CO2 capture.
Owning to their exceptional photophysical and electrical properties, and rigid and tunable structure, porphyrins are frequently incorporated into various porous materials.36–40 The presence of basic pyrrole segments that could facilitate interactions with CO2 renders porphyrin-based POPs attractive potential CO2 adsorbents.36 Combining a porphyrin functional motif with CO2-philic azo groups could lead to porous systems with enhanced CO2 uptake capacity and selectivity. Indeed, reported azo-bridged porphyrin-based POPs exhibit remarkable CO2 uptake capacities up to 3.98 mmol g−1 (at 273 K and 1 bar) and an excellent CO2/N2 selectivity up to 64.3 (at 273 K and 1 bar).41–43 Furthermore, the CO2/N2 selectivity values showed an increase with increasing temperature, which makes the azo-bridged porphyrin-based POPs promising candidates for post-combustion CO2 capture.44 Recently, porphyrin-based azo POPs were also evaluated as catalysts for selective CO2 capture and its conversion to cyclic carbonates.45
In our recent studies, we investigated, experimentally and computationally, structural and functional properties of POPs with different trigonal connectors (triphenylbenzene, triphenyltriazine, triphenylpyridine and triphenylamine) and various nitrogen–nitrogen linkages (azo, azoxy and azodioxy).46–50 The results of the experimental part of those studies emphasized the influence of nitrogen–nitrogen linkages, building units and synthetic routes on the porosity and CO2 adsorption capacities of the synthesized POPs, and identified several promising candidates for CO2 adsorption.47–50
Computational studies facilitate not only the rationalization but also the prediction of the adsorption behaviour of POPs prior to their synthesis. By employing advanced modeling techniques like calculation of binding energies, electrostatic potential values (ESP) and grand-canonical Monte Carlo (GCMC) simulations, we can predict the binding affinities of POPs with CO2, thereby optimizing the design for enhanced adsorption efficiency.46–49 We suggested a simple model to estimate CO2 uptake from the calculated ESP values48 for different stacking modes of 2D layers with hexagonal pores and showed that the introduction of azodioxy in place of azo linkages can promote CO2 binding. Even though this approach mainly relies on highly organized porous architectures, while actual systems are often amorphous, we found it very effective for the fast screening of targeted systems and achieved a nice agreement with experimental data. Pyridine-based azo-linked polymer TPP-azo exhibits CO2 uptake of 32 mg g−1, nicely matching the calculated values determined for configurations with displaced neighbouring layers (31 and 37 mg g−1).48
Herein, we have extended our experimental and computational research on POPs to systems bearing tetragonal tetraphenylporphyrin connector and azo linkages. It has previously been suggested that the introduction of flexible bridges between two phenyl rings (e.g., ether links) in the aromatic diamino monomers used for the synthesis of porphyrin-based azo-linked POPs could induce enhanced porosity compared to e.g., a rigid biphenyl linker.44 To investigate more deeply the effect of flexible diamino linkers on the porosity and CO2 adsorption capacity of azo-linked porphyrin-based POPs (APPs), herein, we synthesized polymers APP-1 and APP-2, with the methylene and ethylene bridge, respectively, between the phenyl rings.
In addition, new porphyrin-based polymers incorporating semi-rigid sulphur (APP-3) and carbonyl bridge (APP-4) were prepared. Besides systems with diamino linkers, we also synthesized polymers APP-5 and APP-6 incorporating trigonal triamino linkers. Like porphyrin-based POPs known from the literature, APP-1 to APP-6 were synthesized by heterocoupling of 5,10,15,20-tetrakis(4-nitrophenyl)-21H,23H-porphyrin (TNPPR) and corresponding aromatic diamines and triamines (Fig. 1). In the present work, we also examined, for the first time to the best of our knowledge, oxidative homocoupling of 5,10,15,20-tetrakis(4-aminophenyl)-21H,23H-porphyrin (TAPPR) for the synthesis of APP-7a and reductive homocoupling of TNPPR and TNPPR-Zn for the synthesis of porphyrin-based polymers APP-7b and APP-8, respectively (Fig. 2). The structural and functional properties of synthesized polymers were thoroughly characterized by IR spectroscopy, 13C CP/MAS NMR spectroscopy, powder X-ray diffraction (PXRD), elemental analysis, thermogravimetric analysis (TGA) and nitrogen (N2) adsorption–desorption experiments. Furthermore, experimental data were augmented by computational study, employing periodic density functional theory (DFT) calculations and GCMC simulations.
![]() | ||
Fig. 1 Synthesis of APP-1 to APP-6 by heterocoupling of TNPPR and corresponding aromatic diamines and triamines. |
The presence of azo groups in APPs was confirmed by FTIR and 13C CP/MAS NMR spectroscopy. The FTIR spectra of all APPs revealed the appearance of bands around 1470 and 1400 cm−1, which can be attributed to the stretching vibrations of the azo (NN) bonds (Fig. 3a–c and S5–S13†). The position of these bands is in good agreement with the stretching vibrations of azo bonds in similar azo-bridged porphyrin-based POPs, which appear at about 1470–1460 cm−1 and 1410–1400 cm−1.43–45,51 In addition, the asymmetric and symmetric stretching bands of unreacted nitro groups around 1510 and 1340 cm−1, respectively, could be observed in the spectra of APPs. In the FTIR spectra of most APPs, the bands around 3400 and 3300 cm−1 assigned to the N–H stretching vibrations of unreacted amino groups were also detected. Nevertheless, the intensity of residual nitro and amino bands in APPs is attenuated when compared to starting nitro and amino monomers. The FTIR spectra of previously reported azo-linked porphyrin-based POPs also showed bands of attenuated intensity located at 1520–1510 cm−1 and 1350–1340 cm−1, and 3500–3200 cm−1, suggesting the presence of residual unreacted nitro and amino groups, respectively.41,43–45,51 The successful formation of azo bridges was finally confirmed by the appearance of the characteristic signal of the carbon directly attached to the azo group (–C–N
N–) around δ = 150 ppm in the 13C CP/MAS NMR spectra of all synthesized APPs (Fig. 4a–c and S22–S30†). The chemical shift of this signal is consistent with that of –C–N
N– carbons in other similar azo-linked POPs.27,29,34,41,43,44,51 Furthermore, signals of carbon atoms of the porphyrin, phenyl and pyridine moieties, mostly situated at around 145, 130 and 118 ppm, were also observed in the 13C CP/MAS NMR spectra of APPs and their chemical shifts are in good agreement with those of previously reported azo-linked POPs.27,29,34,41,43,44,51 In addition, the 13C CP/MAS NMR spectra of APP-1 to APP-6, synthesized by heterocoupling of TNPPR and corresponding diamino and triamino linkers, and APP-7a, prepared by oxidative homocoupling of TAPPR, revealed signal around δ = 118 ppm, which is assigned to the carbon atom near amino group,27 further indicating the presence of unreacted amino end groups.
The elemental analysis revealed some deviations between the experimentally determined and expected composition. Such discrepancies are frequently observed in POPs and can be attributed to incomplete polymerization and the presence of unreacted end groups, as well as adsorption of moisture and gases from air.34,35,52–54
PXRD pattern of all APPs showed broad diffraction peaks, indicating their amorphous nature (Fig. S31†). This is consistent with the previous reports on similar azo-bridged POPs, which are also amorphous due to the irreversible formation of azo bonds.27,29,34,35,41–44,55,56
The thermal stability of APPs was evaluated by TGA. The samples were heated up to 800 °C in N2 atmosphere and held at isothermal conditions for 15 min at 800 °C. Samples APP-1 (Fig. S32a and b†), APP-2 (Figure S33a and b†), APP-3 (Fig. S34a and b†), APP-4 (Fig. S35a and b†), APP-5 (Fig. S36a and b†) and APP-6 (Fig. S37a and b†) display very similar thermal behaviour and show two decomposition events. They are all stable to around 250–270 °C when they slowly begin to decompose and exhibit ≤5.2% initial mass loss in a step that lasts until ∼430 °C when the rate of decomposition increases, and the second decomposition event starts where samples lose an additional 13.0–20.4% of their initial mass (depending on the sample) until approximately 620–640 °C. After that the rate slows down again and a trend of slow decomposition continues until 800 °C where it slows down again during the isothermal step. In all cases the first decomposition event shows slower decomposition rates than the second. Samples APP-7a (Fig. S38a and b†), APP-7b (Fig. S39a and b†), and APP-8 (Fig. S40a and b†) start to decompose at a lower temperature of around 190–200 °C. Sample APP-7a shows similar thermal behaviour as samples APP-1 to APP-6 with a small decrease in mass (<3% of the initial mass) until ∼330 °C when the decomposition rate increases, however, the decomposition occurs in one continuous decomposition event which lasts until ∼650 °C during which the sample loses an additional 30.8% of its initial mass. After ∼670 °C another decomposition event starts and a small increase in the rate is observed which lasts until 800 °C when the rate gradually slows down during the isothermal step. Sample APP-7b shows two decomposition events below 650 °C, however, it is the only sample where the decomposition rate during the first event is higher than during the second event. It begins to slowly decompose (∼2% initial mass loss) until around 280 °C when the rate of decomposition increases, and the sample loses an additional 11% of its initial mass until ∼420 °C. The decomposition continues in the second event above 430 °C and the sample loses 10.4% of its initial mass until ∼640 °C. This sample also shows a third decomposition event starting at ∼720 °C where a small increase in the rate is observed which lasts until 800 °C when the rate decreases again during the isothermal step. The sample APP-8 shows similar thermal behaviour as samples APP-1 to APP-6 with a slight difference in the first decomposition event. APP-8 shows a higher decomposition rate starting at around 240 °C with another increase in the decomposition rate observed at 350 °C resulting in a loss of 6.8% of its initial mass until ∼420 °C. Around 430 °C the second decomposition event starts with an additional increase in the decomposition rate and a loss of 14% of its initial mass until ∼680 °C. The decomposition rate slows down and the slow decrease in mass continues until the end of the isothermal step at 800 °C.
Experimental | Calculated | ||||
---|---|---|---|---|---|
a Different synthetic methods. Labeled as APP-7 in the computational study.b At approximately 306 K.c At a pressure of 1 bar and 298 K. | |||||
Compound | BET surface area (m2 g−1) | CO2 uptakeb (mg g−1) | Average surface area (m2g−1) | CO2 uptakec (mg g−1) | |
APP-1 | 598 | 41 | 1969 | 13 | |
APP-2 | 594 | 38 | 2246 | 18 | |
APP-3 | 586 | 29 | 1931 | 15 | |
APP-4 | 469 | 30 | 1934 | 20 | |
APP-5 | 608 | 38 | 6777 | 49 | |
APP-6 | 500 | 28 | 7567 | 49 | |
APP-7a![]() |
23 | 15 | 1332 | 23 | |
APP-7b![]() |
0.3 | 22 | 1332 | 23 | |
APP-8 | 3 | 28 | 1195 | 21 |
The CO2 adsorption capacities of APPs were investigated by thermogravimetric analysis at approximately 306 K and the data are summarized in Table 1. All measured samples have shown an increase in mass when switching to CO2 atmosphere and a decrease in mass when switching back to N2 atmosphere (Fig. 6a–c and S32c–S40c†). The transition for most samples was sharp (plateau and baseline values reached relatively quickly upon change of atmosphere) and their mass decreased back to baseline values during N2 atmosphere cycles. Samples APP-7a (Fig. 6c) and APP-7b (Fig. S39c†) exhibit distinct behaviour. While most of the mass increase occurred during the first 5 minutes of the CO2 atmosphere cycles, both samples show a gradual increase in mass and APP-7a reaches a plateau after approximately 12 minutes while APP-7b continues to slowly increase in mass during the whole CO2 cycle. These samples also show different CO2 desorption behaviour. While most of the reduction in mass occurs during the first 5 minutes of the N2 cycles the process gradually slows down and baseline values were not reached even after 20 minutes of the N2 cycles. Some of the major factors influencing the CO2 uptake capacity of POPs include their surface area, pore structure and incorporation of heteroatoms in the building units.28 In general, the higher BET surface areas of APP-1 to APP-6, which incorporate various diamino or triamino linkers, led to higher CO2 uptake values compared to APP-7a, APP-7b and APP-8 in which tetraphenlyporphyrin units are directly connected via azo linkages and which show much lower BET surface areas. The highest CO2 adsorption capacities were observed for APP-1 (41 mg g−1) and APP-2 (38 mg g−1) containing flexible methylene and ethylene bridge, respectively, and APP-5 (38 mg g−1) bearing nitrogen-rich melamine linker. These APPs also showed the highest BET surface areas. Nevertheless, as suggested previously, a high value of BET surface area is not necessarily a determining factor for CO2 adsorption affinity, and POPs with BET surface areas lower than 100 m2 g−1 can show good CO2 uptake capacities.27,28,57 This is nicely corroborated from the CO2 adsorption capacities obtained for APP-7a (15 mg g−1), APP-7b (22 mg g−1) and APP-8 (28 mg g−1) which all display low BET surface areas, indicating that the presence of CO2-philic azo groups and pyrrole rings contributes most to CO2 binding affinity of these three APPs. The introduction of Zn to the inner core of the porphyrin ring in APP-8 proves to be beneficial for the increase in CO2 uptake capacity compared to the free base porphyrin polymers APP-7a and APP-7b. Noteworthy, the CO2 uptake capacity of APP-8 of 28 mg g−1 is comparable to the CO2 uptake values of some APPs with diamino (e.g., APP-3 and APP-4) and triamino (e.g., APP-6) linkers, although the BET surface area of APP-8 is much lower. The CO2 uptake capacities of APPs are comparable to some reported azo-linked porphyrin-based POPs such as Azo-CPP-1 (48 mg g−1 at 303 K), Azo-CPP-3 (39.9 mg g−1 at 303 K), Azo-CPP-6 (40.2 mg g−1 at 303 K) and Azo-CPP-7 (41.7 mg g−1 at 303 K).44 In addition, these values are comparable or higher than those of other azo-linked POPs e.g., azo-COPs (15–31 mg g−1 at 323 K) built from 3D tetrahedral building units,29 and pyridine-based TPP-azo (32 mg g−1 at 303 K),48 but lower than those of e.g., benzene-based azo-linked ALP-4 (81 mg g−1 at 298 K).34
![]() | ||
Fig. 6 Representative thermogravimetric CO2 adsorption and desorption profiles of (a) APP-2, (b) APP-6 and (c) APP-7a at approximately 306 K. |
Although, depending on their conformational flexibility, tri- and tetratopic linkers can acquire 2D nets58–60 and various 3D topologies,61 in this paper, we have focused only on topologies commonly referred to as pto topologies.62 Such systems are characterized by low densities and high porosity when compared to other crystalline framework materials.62
The discrepancy between the calculated average surface areas and the experimentally determined values can be attributed to the methodology employed in the computational study. We assumed highly ordered AA structures with perfectly aligned neighbouring layers, as AA eclipsed stacking typically exhibits the lowest energies compared to AB staggered, inclined and serrated configurations.48 However, such highly ordered structures were not achievable experimentally. PXRD experiments indicated the amorphous nature of all investigated APPs, which is reflected in significantly lower BET surface areas.
Overall, systems exhibiting the highest average surface areas, such as APP-5 and APP-6 with pto topologies, also show the greatest available pore volumes, subsequently leading to an enhanced capacity for CO2 adsorption of 49 mg g−1 (Fig. 7e, Table 1 and S7†). The calculated CO2 uptake values are very similar in other APPs with AA stacked layers. Slightly increased CO2 uptakes of 23 and 21 mg g−1 are observed for APPs with directly connected porphyrin rings and rhombohedral pores, APP-7 and APP-8, respectively. In comparison, slightly lower CO2 uptakes of 13 and 15 mg g−1 are found in 2D systems with rectangular-like pores, APP-1 and APP-3, respectively. This indicates that not only the topology and dimensionality of the system (3D vs. 2D) but also pore dimensions and shapes can influence the framework's ability to bind CO2.
While a direct correspondence between calculated and experimental values is not always observed, APP-5, which has the highest surface area, also exhibits one of the highest CO2 uptakes both, experimentally (38 mg g−1) and computationally (49 mg g−1). The experimental and calculated CO2 uptakes are quite similar for the systems with directly connected porphyrin units, such as APP-7 (22 vs. 23 mg g−1) and APP-8 (28 vs. 21 mg g−1). This observation aligns with our previous findings for directly attached trigonal building units,47,48 which partly justified the approach of assuming AA stacking of neighbouring layers. Interestingly, the experimentally observed higher CO2 binding in amorphous APPs with bridged porphyrin units (APP-1 to APP-4) compared to APP-7 and APP-8, which have directly connected porphyrins, shows an opposite trend to their calculated AA analogues.
To further investigate this discrepancy, we generated density plots revealing the spatial distribution of CO2 molecules (Fig. 8a, b and S59†). Regions of high adsorption density are positioned along the pore walls of AA stacked APPs, specifically near the azo linkages and the outer parts of porphyrin rings. Additionally, 3D APP-5 and APP-6 indicated areas above and below the porphyrin rings as highly probable binding sites for CO2 molecules.
Besides the surface area available for CO2 adsorption, the local landscape of the surface, with binding sites of varying shapes and electrostatic potential values (ESP), can also affect the strength of different CO2 binding modes onto the framework's surface. The density plots align well with the calculated ESP maps (Fig. 8a, b and S60†), showing that areas with the most positive or negative ESP values match the spatial distribution of CO2 molecules. The impact of APPs dimensionality on CO2 adsorption is clearly demonstrated by comparing the AA stacked 2D configurations with the 3D networks of APP-5 and APP-6. In all APPs, the inner parts of porphyrin rings display significant negative ESP values, suggesting potential interactions with CO2 molecules. However, such interactions are only achieved if the inner parts of the porphyrin rings have surface area available for binding gas molecules, as seen in the 3D APP-5 and APP-6.
To explore this hypothesis further, we selected APP-2 with linearly bridged porphyrin rings to compare its structural and adsorption properties. This comparison focuses on the effects of neighbouring layer slippage, which opens the inner regions of the porphyrin rings to interact with CO2 molecules. The computational analysis was extended to two limiting cases: the AA (eclipsed) configuration, showing perfectly aligned layers, and the AB (staggered) configuration, with maximally displaced neighbouring layers (Fig. 8c). Although both systems, APP-2AB and APP-2AA, have very similar average surface areas (2388 and 2246 m2g−1, respectively), the unhindered porphyrin ring creates additional binding sites for interaction with CO2 molecules, as confirmed by density plot of APP-2AB (Fig. 8b). This type of structural modification significantly increased the calculated CO2 uptake from 18 to 238 mg g−1 and changed the CO2/N2 selectivity from 4.4 to 21.4. However, APP-2AB is 100.5 kJ mol−1 less stable than its APP-2AA analogue, making it an unlikely configuration to be acquired. As experimentally determined APPs were found to be amorphous, without well-ordered structure, they can potentially have unhindered porphyrin positions. Despite having relatively smaller average surface areas compared to the calculated values of perfectly ordered AA compounds, these amorphous structures can exhibit different CO2 uptake values due to the porphyrin rings being free to interact with CO2 molecules. Nevertheless, the agreement between the 3D (e.g., APP-5 and APP-6) and 2D AA configurations of directly connected building units (e.g., APP-7 and APP-8) is quite good for a preliminary estimate of their CO2 adsorption potential.
The synthesized compounds were identified by solution 1H and 13C NMR spectroscopy, solid-state 13C CP/MAS NMR spectroscopy, IR spectroscopy, PXRD and elemental analysis. Solution-state 1H and 13C NMR spectra were recorded on a Bruker Ascend 400 MHz NMR spectrometer at 25 °C. DMSO-d6 was used as a solvent and TMS as an internal standard for chemical shifts. Solid-state 13C CP/MAS NMR spectra were recorded on a Bruker Avance Neo 400 MHz NMR spectrometer and Bruker Avance III HD 400 MHz NMR spectrometer at spinning rates of 10 and 15 kHz, respectively. FTIR spectra were recorded on a PerkinElmer UATR two spectrometer in the spectral range between 4000 cm−1 and 400 cm−1 at a resolution of 4 cm−1, averaging 10 scans per spectrum. Elemental analysis was provided by the Analytical Services Laboratory of the Ruđer Bošković Institute, Zagreb, Croatia. PXRD diffractograms were recorded on a Malvern Panalytical Aeris powder diffractometer in Bragg–Brentano geometry with PIXcel1D detector. Thermogravimetric analysis was carried out using a simultaneous TGA-DTA analyzer Mettler-Toledo TGA/DSC 3+. Samples were placed in alumina pans (70 μL) and heated in flowing nitrogen (50 mL min−1) from 30 °C up to 180 °C at a rate of 10 °C min−1 and held in isothermal conditions for 10 minutes at 180 °C to remove traces of solvents. Afterwards, samples were cooled to room temperature and heated in flowing nitrogen (50 mL min−1) from 25 °C up to 800 °C at a rate of 10 °C min−1 and held in isothermal conditions for 15 minutes at 800 °C. CO2 sorption experiments were carried out by following a previously reported procedure with minor modifications in experimental conditions.66 Before performing CO2 adsorption experiments, a 70 μL alumina pan was filled with a fresh sample, heated to 100 °C at a heating rate of 20 °C min−1 in nitrogen atmosphere (flow rate 150 mL min−1) and held at 100 °C for 30 minutes to dry the sample. After drying, CO2 adsorption was measured by switching between N2 atmosphere and CO2 atmosphere in 20 min intervals (flow rates for both gases were 150 mL min−1) at ∼30 °C. The measured sample temperature varied around 33 °C during CO2 cycles. To correct for different buoyancy effects on the TG scale and alumina pan, a baseline curve was recorded under the same experimental conditions using an empty alumina pan and subtracted from the measured curve. Data collection and analysis were performed using the program package STARe Software 16.40 MettlerToledo GmbH. The specific surface area was determined from N2 gas adsorption–desorption data obtained with Micromeritics ASAP-2000 at 77 K. Before analysis, samples were degassed at 150 °C under a dynamic vacuum of 7 mPa. The adsorption data were used to calculate the surface area with the BET model, using data points within the linear region of the BET isotherm (P/P0 = 0.05–0.3). The pore size distribution was determined using the Barrett–Joyner–Halenda (BJH) method.
Periodic density functional theory (DFT) calculations were performed in CRYSTAL17 (ref. 71) using the PBE functional72 with Grimme's D3 correction for a better description of the weak dispersive interactions.73 Triple-zeta basis set pob-TZVP-rev2, adapted for periodic calculations, was employed for all atoms.74 The input files for CRYSTAL17 were created from cif files with cif2cell package.75 Full optimization of both atom coordinates and unit cell parameters were performed with default convergence criteria. Total energy convergence was set to 10−7 and truncation criteria for the calculations of coulombs and exchange integrals increased to (8 8 8 8 16) for SCF calculations. The reciprocal space was sampled using 2 × 2 × 8 Pack–Monkhorst k-point mesh for APP-1 to APP-4, APP-7, APP-8 while 2 × 2 × 2 Pack–Monkhorst k-point mesh was used for APP-5 and APP-6. The same level of theory and optimization criteria were used on 2D layers (keyword SLAB for 2D system in CRYSTAL 17) and 3D crystal structures. The lattice parameters and other relevant data of the DFT-optimized structures are given in the ESM.†
The net atomic charges were calculated using the REPEAT method76 from electrostatic potential values calculated by CRYSTAL17. Accessible surface areas of all compounds were calculated by Monte Carlo sampling approach with a spherical probe of the 1.55 Å (nitrogen Wan der Vaals radius) using Zeo++ code.77 The grand canonical Monte Carlo (GCMC) simulation was performed to obtain the single component adsorption isotherms of CO2 at 298 K using RASPA code.78 For the modeling of the interactions between gas molecules and framework, the site–site Lennard-Jones (LJ) potential and coulombic interactions were used together with Lorentz–Berthelot mixing rules for the LJ interactions between different atoms. A three-site model was used to describe the CO2 gas molecules within the TraPPE force field.79 Other atoms from the framework were modeled using the DREIDING force field.80 Default cut-off value, as suggested by the RASPA code, was used for LJ and the short-range part of the coulombic interactions. The long–range interactions were evaluated by the Ewald summation method with a default relative precision of 10−6. The pressure was converted to fugacity using the Peng–Robinson equation of state and further used to calculate the chemical potential. In total 2 × 2 × 8 unit cells were used to obtain a valid GCMC simulation cell with a requirement that all the perpendicular cell lengths have to be larger than twice the default cut-off distance of 12 Å. During the optimisation, framework atoms were frozen. Four different MC moves of gas molecules (translation, rotation, reinsertion, and swap) were allowed during the simulation. More than 106 cycles were used for the equilibration and production phases. The heats of adsorption were also calculated based on the RASPA code, from the limit of using a single adsorbate under zero loading by assuming the framework is rigid. Pore shapes and density pictures for the adsorption of CO2 at 298 K and 1.0 bar were obtained as described in the RASPA manual. Density plots illustrating CO2 binding at 298 K and 1.0 bar were generated using the procedure outlined in the RASPA code and subsequently visualized with the Visualization Toolkit (VTK). The electrostatic potential values are mapped onto the 0.002 a.u. isodensity surface. Additional details can be found in the ESI.†
The computational study supported the experimental findings and provided additional insights into the structural characteristics influencing CO2 uptake. Despite the computational models being based on ideal structures while the experiments revealed the materials were amorphous, the calculated CO2 adsorption capacities were roughly comparable to the experimental results, particularly for the 3D systems (e.g., APP-5 and APP-6) and the 2D systems with directly connected building units (e.g., APP-7 and APP-8). Density plots and electrostatic potential maps highlighted the importance of available surface areas and their local landscape on CO2 adsorption. The porphyrin building units introduced into the framework act as additional binding sites for CO2 molecules, especially when the units are unhindered and available to bind CO2 molecules on either side of the porphyrin plane.
Finally, this work represents the successful construction of APPs through the rational design of carefully selected connectors and linkers. The experimental study indicated that 2D APP analogues with linkers and 3D topologies may enhance CO2 adsorption performance compared to directly attached tetraphenylporphyrin units. This advancement highlights the potential applications of APPs as efficient CO2 adsorbents.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra08113g |
This journal is © The Royal Society of Chemistry 2025 |