Next Article in Journal
Enterprise Spatial Agglomeration and Economic Growth in Northeast China: Policy Implications for Uneven to Sustainable Development
Previous Article in Journal
Sustainable Management of Marine Protected Areas in the High Seas: From Regional Treaties to a Global New Agreement on Biodiversity in Areas beyond National Jurisdiction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of CO2, CO, H2, and N2 on Zeolites, Activated Carbons, and Metal-Organic Frameworks with Different Surface Nonuniformities

Department of Environmental Engineering, Daegu University, 201 Daegudae-ro, Jillyang, Gyeongsan 38453, Republic of Korea
*
Author to whom correspondence should be addressed.
Sustainability 2023, 15(15), 11574; https://doi.org/10.3390/su151511574
Submission received: 12 June 2023 / Revised: 7 July 2023 / Accepted: 12 July 2023 / Published: 26 July 2023

Abstract

:
The single-component adsorption of CO2, CO, N2, and H2 at 25 and 35 °C was studied using microporous faujasite-framework zeolites (NaY and NaX), activated carbons (GCN and MSP), and metal–organic frameworks (A100 and Z1200) as starting points for the separation of CO2 from syngases produced by gasifying biomass-based solid wastes. The indicated adsorption isotherms and uptake of the adsorbates strongly depended on the adsorbates themselves as well as on the adsorbents because of significant differences in the surface features, such as surface nonuniformity, and in the molecular properties. The selectivity of CO2 to the other gases also varied with the adsorbents due to the distinctive energetic characteristics. The surfaces of the zeolites were the most energetically heterogeneous ones, yielding higher CO2 uptake at low pressures, while the two activated carbons and A100 had moderate surface heterogeneities, and MSP showed the highest CO2 uptake at high pressures, such as 6 bar, at which the micropore volume and surface area are important. Z1200, which has highly homogeneous surfaces and no high-affinity-binding sites, exhibited the lowest CO2 adsorption capacity regardless of equilibrated pressure. The surface nonuniformities of the six sorbents were consistent with the calculated isosteric heats of CO2 adsorption. CO2 could be reversibly adsorbed on NaY and MSP but not on GCN, with some metal impurities, although all these adsorbents showed a fully reversible process for CO adsorption. The estimated working capacity for CO2 adsorption at 25 °C was 0.78–6.50 mmol/g, depending on the sorbents used. The highest value was disclosed for MSP, the surface energetic heterogeneity of which was between that of zeolites and Z1200. Such a high working capacity bodes well for use in our later applications.

1. Introduction

Gasification of sustainable resources, such as biomass (e.g., wood and wood processing wastes and forest and agricultural residues), can yield syngases that consist of H2, CO, CO2, and N2, with other gaseous components in relatively small quantities, and many kinds of sustainable biofuels and chemicals can be driven from such syngases [1,2]. For this, it is necessary to separate undesirable gases, such as CO2 and N2, that exist with H2 and CO from the main stream using appropriate processes [3,4]. Numerous studies have been conducted on adsorptive CO2 separation that combine a mechanism to preferentially adsorb the guest in a mixture with other gases on the surfaces of porous sorbents with a way to recover their bare surfaces, such as pressure-swing adsorption (PSA), thereby allowing for the periodic adsorption–regeneration operation [4,5,6,7]. To date, numerous porous solids have been proposed as promising CO2 sorbents [6,8,9,10,11,12,13].
Zeolites are favored for the adsorption of gaseous adsorbates because of the presence of cationic sites [4,5]. Among them, low-silica zeolites, such as Y, 13X, and A, are preferred for separating CO2 from gas mixtures with CO, N2, H2, CH4, and O2, owing to their high affinity toward the target gas [6,14], although such low-SAR (Si-to-Al ratio) zeolites often have an issue that complete regeneration using a pressure swing technique is difficult [15]. Amorphous activated carbons (ACs) with high micro- and mesoporosity have been frequently studied as representative carbon-based materials for CO2 separation [4,7,16], while molecular-sieve-type carbons reported for CO2/CH4 separation from landfill gases [17] are relatively more expensive than ACs, making them inappropriate for widespread use in the gas separation industry. Such ACs contain many types of surface functional groups (e.g., carbonyl, phenol, carboxyl, lactone, and ether) that have an acidic property but that can be altered to basic centers after thermal treatment, making them good binding sites for CO2 [18,19]. Not only do these surface properties depend dramatically on the carbon sources and activation processes, but the impurity metals present in ACs also have a substantial influence on the adsorption of adsorbates [4,19,20]; thus, it may not be easy to develop AC sorbents with on-demand physical and surface properties for CO2 separation, such as a well-defined pore size and distribution, channel system, and surface functionality, unlike zeolites. Since the first synthesis of CoC6H3(COOH1/3)3 with a three-dimensional network of channel and uniform pores by Yaghi et al. [21], the use of other sources for metal-based organic building blocks and organic linkers has led to tremendous crystalline metal–organic frameworks (MOFs), including zeolitic imidazolate frameworks (ZIFs) [9,10,12,13,22,23,24,25], enabling their framework structures and surface properties to be rationally designed and tuned, which makes them different from the benchmark porous solids, such as zeolites and ACs. Some MOFs possess not only high BET surface areas (SBET) compared to zeolites and ACs (e.g., 4230 for MIL-101 [26], 4500 for MOF-177 [27], 5100 for PCN-68 [28], and 6240 m2/g for MOF-210 [29]) but also extremely high-accessible pore volumes (e.g., MOF-200 (SBET = 4530 m2/g) and MOF-210 with ca. 3.59 cm3/g) [29]. One of the most intriguing properties of many MOFs is that the metal centers after quest removal can become accommodation sites for adsorbates, similar to zeolites. Due to such textural and surface features, MOFs are extensively studied for the adsorptive separation of desired gases, such as CO2, from gas mixtures containing other light gases [12,13,25,29,30].
Most of the early studies on adsorptive gas separation using zeolites, ACs, and MOFs have predominantly focused on CO2/CH4 systems [31,32,33] because of the substantial market need to purify natural gas, which contains 80–95% CH4 and 5–10% CO2, with N2 and heavier hydrocarbons making up the remainder of the gas [20]. The suitability of such porous solid sorbents may depend on their potential applications. Exceptionally high CO2 uptake (30.4–54.5 mmol/g) has been reported on recently spotlighted MOFs, such as MOF-200 and MOF-210, MIL-101, PCN-68, and MOF-177 [28,29,30,34], at 25 or 31 °C. However, such CO2 adsorption has been possible at pressures near 35–50 bar (1 bar = 105 Pa), which is too high because many industrial PSA systems commonly allow adsorption at 4–6 bar [14,20]. To the best of our knowledge, few early studies were able to directly compare the performances in CO2 adsorption between the traditional benchmarks and the new class of materials at the typical PSA pressure. Indeed, Garces et al. [35] established a comparison of CO2 adsorption on such solid adsorbents, but they allowed the use of only up to 1.2 bar pressure, which is almost the pressure for desorption (regeneration) rather than adsorption in typical PSA processes. We are, of course, aware of the adsorption of CO2 and other gases, such as CH4, N2, H2, CO, and N2O, on two different types of porous sorbents, i.e., zeolites (13X, 4A) and ACs [36,37,38], and zeolites (13X, 5A) and MOFs (Cu-BTC, MOF-5, MOF-177) [39,40]. However, all these studies were conducted at adsorption pressures that were too low (slightly over the atmospheric pressure).
With the goal of producing sustainable syngases from biomass-based industrial wastes, their gasification under well-controlled conditions is under investigation in another part of this research program. As per our preliminary results, the gas streams produced were composed of approximately 20% CO2, 38% CO, 4% N2, and 38% H2, with gases such as HCN, NH3, HCl, and H2S at tens to hundreds of ppm levels, varying with raw biomass wastes fed to the gasifier and its operation conditions. Thus, it is essential to primarily separate the 20% CO2 from the mixture using cost-effective, energy-efficient separation technologies, such as PSA. For this, we studied the pure-component adsorption of CO2, CO, N2, and H2 on zeolites, ACs, and MOFs up to a maximum pressure of 7.5 bar and at temperatures of 25 and 35 °C, conditions that are similar to typical operating conditions (T = 15–40 °C; P = 4–6 bar) for many industrial PSA systems [14,20,41]. The porous adsorbents chosen here have a significant difference in surface energetic heterogeneity, and the adsorbates possess distinct molecular properties, such as polarizability (α) and quadrupole moment (qm). Though our ultimate goal was to separate CO2 from the gas mixture, it should be emphasized that this study was conducted using only a single component of each gas in the specified range of temperature and pressure to examine the influence of surface nonuniformity on the molecular adsorption of light gases.

2. Materials and Methods

2.1. Adsorbents

NaY (CBV 100, Zeolyst, Conshohocken, PA, USA) and NaX (Z10-08EP, Zeochem, Houston, TX, USA) were used as representatives of zeolites. Two kinds of ACs were used for this study: a Norit GCN612 produced from coconut shells using a steam activation process, hereafter designated as “GCN”, and an MSP-20 phenolic-resin-based AC (Kansai Coke and Chemicals, Amagasaki, Japan), denoted as “MSP”. MOFs chosen here were Basolite A100 and Basolite Z1200 (Aldrich, St. Louis, MO, USA). The former is an aluminum-based MOF isostructural to MIL-53(Al) [9], and the latter is a Zn(II)-based ZIF with a zeolite-like topology that is an isostructural form of ZIF-8 synthesized by Yaghi and coworkers [10]. These powder-type MOF analogues are, respectively, referred to as “A100” and “Z1200”.

2.2. Characterization of Microporous Adsorbents

Liquid N2 isotherms of the adsorbents used were collected using a Micromeritics ASAP 2020. A small amount of each sample (ca. 40 mg) was loaded into a factory-made adsorption cell in the apparatus and evacuated for 1 h at 30 °C, for 1 h at 90 °C, and finally overnight at 200 °C in vacuum, before N2 was introduced into the sample cell at −196 °C. SBET values were determined according to the BET multipoint technique. The size of the micropores present in the evacuated samples was calculated from the N2 sorption data using the original Horvath–Kawazoe model, assuming slit-type micropores for ACs, and the Saito–Foley cylindrical pore model for the others [42,43], while porosity data regarding mesopores were obtained using the Barrett-Joyner-Halenda (BJH) pore model. Finally, quantities of Al and Si in zeolite samples were determined using inductively coupled plasma (ICP) measurements, and metal impurities existing in AC sorbents were qualitatively acquired using a Rigaku Model ZSX100e wavelength-dispersive X-ray florescence (XRF) spectrometer.

2.3. Volumetric Adsorption Measurements

The adsorption of CO2, CO, N2, and H2 on the adsorbents chosen was carried out using a stainless steel high-vacuum system with a dynamic vacuum below 10−8 Torr (1 Torr = 133.3 Pa), equipped with an Oerlikon Leybold Vacuum Model TURBOVAC 151 turbomolecular pump (TMP) with a Turbo.Drive TD 20 Classic controller backed with a Kodivac GHP-340K mechanical pump. Dynamic vacuum levels were monitored with a Pfeiffer Model TPG 261 single-gauge controller. During the measurements, changes in gas pressure were measured using an absolute Honeywell Model Super TJE ultra-precision pressure transducer (Type AP112) with a full range of 200 psi (1 psi = 6.896 × 103 Pa) connected to a Sensys Model 1300 pressure indicator, and the temperature of a manifold of the system was monitored with a Hanyoung Nux Model BK-6M Digicator. A schematic of such an adsorption system is shown in Figure S1, and details of a similar volumetric system have been also provided elsewhere [44,45,46].
An appropriate amount of each adsorbent (ca. 0.5 g) was placed over a 1 μm stainless steel filter welded inside a stainless steel adsorption cell in a cylindrical electric furnace with a Misung S&I Model TC500P temperature controller, heated to 300 °C using a heating rate of 3 °C/min under relatively low vacuum (10−3 Torr) with a rotary vane pump and evacuated overnight, cooled to room temperature, and further treated under high vacuum (typically 10−7 Torr) with the help of the TMP. After this pretreatment, the electric furnace was replaced by a Peltier-type thermoelectric heating/cooling device coupled with an Omega Model CN 7500 temperature controller so the adsorption cell could be constantly controlled at either 25 or 35 °C, at which adsorption measurements were allowed. The time to be equilibrated at every dose did not exceed 2 min, irrespective of the adsorbent and the adsorption temperature and pressure. Temperatures of the adsorption cell during sample pretreatment and gas adsorption were monitored using an Omega Model 410B Digicator with a K-type thermocouple that was inserted just above the powder sample in the cell through a feedthrough. Before CO2 (99.999%), CO (99.998%), N2 (99.9999%), and H2 (99.9999%, Linde, Danbury, CT, USA) were admitted into the adsorption cell, all the adsorbates were further purified by allowing them to flow through Alltech moisture traps and Oxytraps.

3. Results

N2 adsorption–desorption measurements with NaY, NaX, GCN, MSP, A100, and Z1200 were conducted using the standard BET technique (Figure S2, see Supplementary Materials). All the samples basically showed the standard International Union for Pure and Applied Chemistry (IUPAC) type I isotherm [47], which is a typical characteristic of microporous framework substances. The N2 sorption data were subjected to the standard BET porosity analysis protocol with the help of appropriate pore models for collecting their textural properties, such as SBET, micropore opening (dmc), micropore volume (Vmc), mesopore volume (Vms), and total pore volume (Vt). All these results, including the SAR values of the two zeolites, are listed in Table 1.
SBET values were, respectively, 847 and 724 m2/g for NaY and NaX samples. Both these zeolites with dmc = 7.71 ± 0.13 Ǻ had a Vmc of 0.292 ± 0.021 and a Vt of 0.345 ± 0.011 cm3/g, which was in reasonable agreement with earlier reports on such faujasite-framework zeolites [48,49]. GCN had an SBET of 1132 m2/g, while MSP had an SBET of 2508 m2/g, which was the largest among all porous sorbents used. The micropore sizes of these two carbonaceous substances (respectively, 5.02 and 5.53 Ǻ) could be considered almost the same; however, each corresponding pore volume was significantly different (Table 1). The textural properties of A100 could be characterized to be SBET = 838 m2/g, dmc = 10.96 Ǻ, Vmc = 0.361 cm3/g, and Vt = 1.109 cm3/g. The indicated SBET and Vmc values were in excellent accordance with those (830 m2/g and 0.358 cm3/g) reported for commercial A100 in the form of a powder [50]. However, the dmc and Vt values in our measurements differed from those in the literature. Heymans et al. [50] reported a Vt of 0.549 cm3/g, which is comparable or close to the values (respectively, 0.508 and 0.581 cm3/g) for laboratory-synthesized MIL-53(Al) [51] and another commercial A100 sample [52], while Fierro and coworkers [53] reported a dmc of 5.3 Ǻ and a BJH pore volume of 5.5 cm3/g (thus accounting for only Vms) for a commercial powder sample of A100. Such differences might be associated with the well-known large breathing phenomenon [54]. Pores having dmc = 12.27 Ǻ, Vmc = 0.553 cm3/g, and Vt = 0.739 cm3/g existed in Z1200, with SBET = 1301 m2/g, in reasonable agreement with the values (dmc = 11 Ǻ, Vmc = 0.67 cm3/g, Vt = 0.77 cm3/g, and SBET = 1417 m2/g) reported for a sample of Basolite Z1200 [55]. Finally, the well-known t-plot method gave Vmc values consistent with those compiled in Table 1 (not included here).
The adsorption of CO2, CO, N2, and H2 on the six adsorbents at 25 °C was measured as a function of equilibrated pressure up to ca. 7.5 bar. Figure 1 and Figure 2 show the isotherms for the adsorption of each adsorbate on the respective NaY and NaX. NaY had much greater CO2 uptake at all pressures covered compared to the other gases. At 6 bar, the extent of the adsorption of the gases was CO2 (6.70) > CO (1.95) > N2 (1.08) > H2 (0.13 mmol/g). The type I isotherm was visibly shown for CO2 but not for CO, N2, and H2. The adsorption of H2 on NaY obeyed Henry’s law at all pressures because its uptake passed through zero when the isotherm was extrapolated to zero pressure, while the isotherms for CO2, CO, and N2 obeyed the law only below low or medium pressure, depending on the adsorbates. At high pressures (>3.5 bar), NaX gave a CO2 uptake of nearly 5.4 mmol/g (Figure 2), which is lower, by 1.3 mmol/g, than that measured for NaY. When CO and N2 were adsorbed on NaX, their uptake was higher than that obtained for NaY. There was no difference in the extent of H2 adsorption between both zeolites at all given pressures.
Isotherms for the adsorption of CO2, CO, N2, and H2 at 25 °C on the two kinds of ACs, GCN and MSP, are shown in Figure 3 and Figure 4, respectively. Both carbonaceous adsorbents showed not only a similar shape of each corresponding gas isotherm but also a comparable capacity to adsorb CO, N2, and H2. They, however, had a large difference in the CO2 uptake at high pressures; for example, at 6 bar, 6.13 mmol CO2/g could be adsorbed on GCN, while 10.41 mmol CO2/g was adsorbed on MSP, which is almost 2-fold greater. However, the isotherms of CO, N2, and H2 with the AC sorbents were similar to those for the zeolites, whereas the CO2 isotherm differed from that on the zeolites. Such a difference in CO2 adsorption between the ACs and zeolites could be related to surface nonuniformity with high-affinity binding sites for CO2 [12,17].
The extent of the adsorption of CO2, CO, N2, and H2 at 25 °C on A100 and Z1200 after evacuation at 300 °C is shown in Figure 5 and Figure 6, respectively. At all pressures, on both adsorbents, the adsorption of CO2 was the largest, followed by the adsorption of CO, N2, and H2, in that order, which was the same as what appeared for the zeolites and ACs. A100 consistently yielded a higher CO2 adsorption capacity than that obtained with Z1200 at all pressures, suggesting that the former’s surfaces could be somewhat more electrophilic [50,53,56]. It is noteworthy that the CO2 adsorption isotherm of A100 was similar to that of the AC adsorbents. On Z1200 surfaces, even CO2 adsorption was seen at whole pressures that was almost governed by Henry’s law. This suggests that all the adsorbates interacted weakly with this sorbent.
Key parameters that must be assessed in gas adsorption studies for separation are the selectivity factor and the working capacity for a target gas, primarily CO2 here, and these can be estimated using the adsorption data shown before. For multicomponent gas mixtures, the selectivity factor is usually calculated using the correlation (ni/pi)/(nj/pj), where ni and nj are the adsorbed amounts of components i and j, respectively, on the adsorbent surface and pi and pj are the mole fractions of components i and j, respectively, in the bulk gas phase [4,7,20,57]. However, since isotherms of only a single component were collected in this study, we calculated the selectivity factor using a simple correlation (ni/nj) proposed by Corma and coworkers [14], where ni and nj represent, respectively, the molar uptake of gasses i and j at a given pressure in the corresponding pure-component isotherms. For this, we compiled the molar uptake values of CO2, CO, N2, and H2 adsorbed on each adsorbent from the isotherms in Figure 1, Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6, as partly listed in Table 2, in which such uptake values at 1 and 6 bar were collected because of a common operating pressure window for adsorption–desorption cycles in most PSA systems [20]. Figure 7 shows the CO2/CO, CO2/N2, and CO2/H2 selectivities of all the adsorbents with respect to equilibrated pressure. All the estimated selectivities decreased with equilibrated pressure, regardless of the adsorbents used, which is in excellent agreement with previous reports [14,37].
The only exception was Z1200, in which all the selectivities somewhat increased with pressure up to ca. 2.5 bar and then were almost constant, although such pressure dependency on selectivity varied somewhat with adsorbates. To allow a ready comparison, the selectivities for the six porous materials at 1 and 6 bar are listed in Table 2. Finally, the CO2 working capacity is defined as the difference between specific equilibrium molar uptake values at high and low pressures that were, respectively, chosen to be 6 and 1 bar here because of common adsorption–desorption cycles for the PSA processes [14,20]. The estimated CO2 working capacities for each adsorbent are also collated in Table 2.
It is imperative to explore whether the adsorbents used in this adsorption study can be easily regenerated using appropriate desorption techniques. Three repeated measurements for CO2 adsorption on NaY, GCN, and MSP following each evacuation at 25 °C for 2 h are provided in Figure 8, in which the first adsorption isotherms were measured on the clean surface of each adsorbent. These consecutive CO2 adsorptions on NaY and MSP surfaces were completely in line with the first measurement isotherm corresponding to each sorbent, as shown in Figure 8a,c, indicating that CO2 adsorption could be fully reversible and their surfaces could be easily recovered after evacuation, in good agreement with results reported for CO2 adsorption on a zeolite 13X and commercial G-32 H activated carbon [37]. Obviously, there is a limit to these measurements, owing to the use of evacuation under a 10−8-Torr dynamic vacuum, in addition to a pressure swing as a driving force to desorb the adsorbed CO2. Furthermore, there are reports that complete regeneration of zeolites after CO2 adsorption is improbable [11,37]. To resolve this point, we desorbed CO2 from NaY that had undergone CO2 adsorption for the third time using a fast pressure swing from nearly 7.5 bar to 1 bar within 1 min following CO2 adsorption, as shown in Figure 8a. The CO2 isotherm measured from 1 bar was identical to isotherms indicated for the evacuated surfaces, once again showing that CO2 adsorption on the zeolite NaY is a fully reversible process. Meanwhile, after the first exposure to CO2 at 25 °C and up to 7.5 bar, there was irreversible CO2 uptake on GCN at pressure > 2.5 bar, as shown in Figure 8b. However, at whole pressures, fully reversible CO2 adsorption occurred on a carbon surface with small CO2 residues, as revealed by the last isotherm. As for the NaY, both ACs were completely regenerated under the fast pressure swing to 1 bar after each third isotherm measurement (not shown here).
We were interested in metal impurities existing in GCN activated carbon to find out why small amounts of CO2 were irreversibly accommodated on a clean surface of this sorbent in a high-pressure region. The wavelength-dispersive XRF spectrum of an as-received GCN sample was collected to identify such metal impurities and compared to that measured for MSP that had shown no irreversible CO2 adsorption, and these XRF spectra are shown in Figure 9. The original intensity of a peak at 2θ = 136.55° due to K present in GCN reduced by 90% for a ready comparison with peaks of other elements, as indicated by an asterisk in Figure 9a. The most striking difference between the two AC samples was the 136.55° peak by K, the intensity of which was much greater in the XRF spectrum for GCN. Both ACs indicated the presence of transition metals, such as Zn, Cu, Ni, and Fe, but all these impurities were relatively at low levels. Peaks by Mg, Na, Si, Ca, and Al were observed with GCN but not or weakly observed with MSP; however, the effect of such impurities on irreversible CO2 adsorption might be negligible compared to the effect of K.
Isotherms for CO adsorbed repeatedly on NaY, GCN, and MSP are provided in Figure 10. All samples showed that such repeated CO adsorption after every evacuation was on a single isotherm, which implies that there is no irreversible CO adsorption. This indicates that unlike the previous repeated CO2 adsorption on a sample of GCN, the metal impurities, particularly K, in this adsorbent played no appreciable role in the CO adsorption. To confirm that CO adsorption is fully reversible even under a simulated PSA condition, after the third CO adsorption, NaY was pressure-swung to 1 bar in a way similar to that described for CO2 adsorption, and then CO was again introduced into the adsorption cell. This result is shown in Figure 10a, as indicated by the open symbol, and we found that fully reversible CO adsorption occurred on NaY surfaces. After a similar pressure swing, all the ACs also underwent fully reversible CO adsorption (not shown here). It is clear that all these adsorbents are easily regenerated for CO adsorption even when such a pressure swing technique is used.
Isosteric heats of adsorption of CO2 on the six porous materials used were calculated by applying the Clausius–Clapeyron equation,
Δ H =   R T 2 l n P T n = R l n P 1 T n ,
where ΔH is the isosteric heat of adsorption (kJ/mol), R is the universal gas constant, T is the adsorption temperature (K), P is the adsorption pressure (bar), and n is the adsorbed CO2 amount (mmol/g). To do this, high-resolution CO2 isotherms on NaY and NaX at the three temperatures were collected with a narrow pressure interval because of steep changes in CO2 adsorption with these sorbents at low pressures, while for the other adsorbents, CO2 isotherms at 35 and 40 °C were measured additionally to make corresponding sets with the 25 °C adsorption data in Figure 3, Figure 4, Figure 5 and Figure 6. Details of the fits and parameters of the widely used Toth model [58] and the ΔH calculations are provided in Figure S3 and Table S1. Changes in ΔH were plotted as a function of CO2 uptake, as shown in Figure 11. At coverage near zero, the ΔH values on NaY and NaX were 40.8 and 53.8 kJ/mol, respectively (Figure 11), which are similar to those reported in the literature [59,60], and these values decreased up to loadings of ca. 6.1 and 3.9 mmol/g, respectively, after which, both values increased. Samples of GCN and MSP had almost the same ΔH values (25.4 and 23.3 kJ/mol) at zero loading, which are similar to the adsorption energies reported for other kinds of ACs [16,38]. However, these two ACs exhibited different behaviors in the decrease in ΔH with the CO2 amount adsorbed; that is, the ΔH values rapidly decreased on GCN but not on MSP, suggesting that the former surfaces are much more heterogeneous, perhaps due to significant levels of impurity metals, such as K, as disclosed by the XRF measurements in Figure 9. A zero-coverage ΔH of 17.4 kJ/mol was indicated for Z1200, and this value remained constant. The indicated value is similar to that (15.9 kJ/mol) computed for CO2 adsorption on ZIF-8 using a force field optimization simulation technique [61]. The value of ΔH with A100 was around 26.3 kJ/mol, comparable to that (20.1 kJ/mol) reported for MIL-53(Al) [62] but significantly lower than that (63 kJ/mol) determined calorimetrically for CO2 adsorption on MIL-100 at 30 °C [34].

4. Discussion

4.1. Effect of the Textural Properties of Adsorbents on the Adsorption of Adsorbates

The range of the indicated micropore volume fraction, fmc, was ca. 0.75–0.94, except for A100, in which Vms was dominant. The smallest mouth size of micropores was 5.02 Ǻ, which was obtained for GCN. The kinetic molecular diameters of CO2, CO, N2, and H2 used as adsorbates are, respectively, 3.3, 3.69, 3.64–3.80, and 2.83–2.89 Ǻ [12,63]. Thus, it is beyond doubt that no molecular sieving effect is anticipated in the adsorption of the gases even on NaX, because its dmc value is much larger than that of the biggest gas molecule (N2). It is noteworthy that all the isotherms were collected not only by adsorbing only a single gas component on each sorbent but also by allowing an adequate equilibration time upon every measurement; therefore, rate-dependent adsorption, known as the kinetic effect [4,5], is also improbable. In such an adsorption situation, the equilibrium-dependent mechanism would be predominant.
There was a visible difference in SBET values and microporosity between the six adsorbents, as shown in Table 1. The question is whether such differences could contribute to the distinctive uptake of each adsorbate, particularly CO2, among the sorbents. The SBET values were in the following order: MSP >> Z1200 > GCN > NaY ≈ A100 > NaX. The order of CO2 adsorption was NaY (5.60) > NaX (4.60) > MSP (3.25) > GCN (2.75) > A100 (2.20) >> Z1200 (0.66 mmol/g) at 1 bar and MSP (9.75) >> NaY (6.70) > GCN (5.87) > NaX (5.38) > A100 (5.18) > Z1200 (3.95 mmol/g) at 6 bar (see Table 2). Thus, we could find no consistency between SBET and the CO2 adsorption capacity at either pressure. The Vmc calculations were in the order: MSP >> Z1200 > GCN > A100 > NaY > NaX (Table 1), basically similar to that of the SBET values. However, Z1200 with relatively high SBET and Vmc values yielded the lowest CO2 uptake at both pressures. Consequently, textural data, such as SBET and Vmc, could not be successfully correlated with the indicated uptake of the adsorbates.

4.2. Molecular Properties of Adsorbates and Their Adsorption Behaviors

The adsorbates chosen possess different molecular properties, such as α and multipole moments [4,5,17]. The α value of CO2 is 26.5–29.11 × 10−25 cm3, which is the largest among the adsorbates, and CO, N2, and H2 have α values of 19.5 × 10−25 cm3, 17.403–17.6 × 10−25 cm3, and 8.0–8.042 × 10−25 cm3, respectively [12,17]. The dipole moment of CO is 0.1098–0.112 × 10−18 esu·cm, where esu is the electrostatic unit (1 esu = 3.336 × 10−10 C); however, all the others are non-polar molecules and have no permanent dipole moment, thereby having no electrostatic dipole–adsorbent interactions [4,12,17]. The qm values for CO2, CO, N2, and H2 are 43.0, 25.0, 15.2, and 6.62 × 1025 esu·cm2, respectively [12,17]. Thus, CO2 also has the largest qm value, which is greater, by 1.5–6.5 times, than the qm values of the other gases. A combination of such molecular properties of adsorbates and textural and surface features of adsorbents shows a general trend—the higher the SBET of porous adsorbents, the larger the adsorption of gas molecules with a high α value, while an adsorbent surface with a high electric field gradient would prefer adsorbates with a high qm value [4,5].
Keeping the previous discussion in mind, the capability of the three categorized sorbents (zeolite, AC, and MOF) to accommodate the adsorbates at the conditions chosen, i.e., 25 or 35 °C and up to 7.5 bar, was of particular interest because of the potential candidate to selectively separate CO2 and N2, primarily the former component, at this stage, after the gasification of biomass solid wastes. At all the pressures, the measured uptake of the adsorbates gave consistent results with their molecular properties. The six adsorbents all displayed equilibrium adsorption amounts along the series CO2 > CO > N2 > H2 (see Figure 1, Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6 and Table 2), which is equal to the order of the α and qm values, although their absolute uptake depends on the sorbents. This implies that the larger the α and qm values of a gas molecule, the greater the extent of its adsorption on porous solid sorbents. However, the qm value of CO2 might not contribute much to its high adsorption on MSP-20 at 6 bar. The reason is the almost negligible electrostatic adsorbate–adsorbent interactions, since charges existing on such ACs are weak or close to each other and it is difficult for them to exert a significant electric field or field gradient on the surface [4]. Hence, the van der Waals interactions via dispersion and repulsion forces, whose equations are a linear function of α but not of qm, play a dominant role in the adsorption of adsorbates on ACs. Based on the previous discussion of the textural and surface properties of the adsorbents and the electronic properties of the adsorbates, it is clear that both α and qm properties are strongly associated with the extent of these adsorptions on those materials, except for AC sorbents, on which the highest CO2 uptake mainly results from the largest SBET value, indicating that porous solids with a high electric field gradient, thereby creating strong quadrupole–adsorbent interactions, may be better for the adsorption of CO2 with a high qm value. Such interactions take place on adsorbents with a variety of sites that can have a wide range of adsorption affinities [4,12,36,37] and influence not only the shape of the isotherms of adsorbates, particularly CO2, but also the selectivity of CO2 to other gases, which will be thoroughly discussed next.

4.3. Surface Nonuniformity of Adsorbents and Their Selectivity of CO2 Adsorption

We can obtain the difference in surface properties among the six adsorbents from the shape of the adsorption isotherms for CO2, CO, N2, and H2 at 25 °C. In the CO2 adsorption data shown in Figure 1, Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6, two zeolite samples, i.e., NaY and NaX, exhibited an initial steep rise in its uptake within 1 bar, which differs from the isothermal behavior of the other sorbents. This is characteristic of porous solid substances with high-affinity sites on which CO2 is strongly adsorbed [36,37]. In contrast, CO2 uptake on Z1200 linearly increased with equilibrated pressure (Figure 6), suggesting that this sorbent predominantly consists of CO2 adsorption sites with an affinity weaker than that of even MSP-20 and GCN. These ACs and A100 showed a CO2 isotherm that could be characterized by an intermediate behavior between zeolites and Z1200 (Figure 3, Figure 4 and Figure 5). A similar trend was, to a much lesser extent, observed in the isotherms for CO and N2, although it was not appreciable for H2 because of the lowest α and qm values.
The two zeolites possess extra-framework sites (Na+) well known to be occupied at positions with four- and six-membered Al–O–Si rings [49,64]. The surfaces even with only a single metal species are energetically nonuniform because of such different coordination environments, and the cations on the surface would create an electric field and field gradient that facilitate the adsorption of gas molecules with high multipole moments due to the extra-framework charge–multipole interactions [65]. CO2 has the greatest qm value (but no dipole), as discussed previously, and would undergo strong charge–quadrupole interactions; thus, the zeolites yield the steepest increase in CO2 adsorption at low pressures (Figure 1 and Figure 2). However, ZIF-8 has been known to contain no coordinatively unsaturated sites (CUSs) interacting with adsorbates [61], and Z1200 isostructural to ZIF-8 exhibits isotherms with surface characteristics of the weakest affinity to even CO2. In the case of GCN, MSP, and A100, the intermediate interaction between their surfaces and gas molecules is due to the presence of moderate-strength affinity sites. Appropriate thermal pretreatments of carbonaceous materials with various oxygen-containing functional groups result in their conversion to heterogeneous surfaces with a highly basic nature [66], which is preferable to CO2 adsorption. However, even after the ACs studied here were pretreated at 300 °C in high-dynamic vacuum overnight, their electrostatic affinity to CO2 seemed much weaker than that of the zeolites but not weaker than that of the sorbent without CUSs, i.e., Z1200. A100 as an isostructural form of MIL-53(Al) consists of chains of corner-sharing AlO4(OH)2 octahedra interconnected by benzenedicarboxylates [9]. The OH groups, denoted as μ2-OH [56], are not a CUS (open metal sites); regardless, it has been frequently proposed that they can somewhat strongly interact with gas molecules with high α and qm properties, such as CO2 [32,56]. Consequently, the shape of the isotherms measured for the adsorbates, particularly CO2, is a good indicator of high-affinity sites in the sorbents, and the indicated order of the surface energetic heterogeneity is zeolites >> ACs ≈ A100 > Z1200.
The previous discussion of the surface nonuniformity of the adsorbents is readily supported by changes in the ΔH value calculated as a function of CO2 coverage. At almost zero loading, the order of the ΔH value was NaX (53.8) > NaY (40.8) >> A100 (26.3) ≈ GCN (25.4) ≈ MSP (23.3) > Z1200 (17.5 kJ/mol), as shown in Figure 11. The zeolites showed the steepest decrease in ΔH values with CO2 loadings, indicating the highest surface heterogeneity with a wide range of adsorption sites with different activation energies [4,8,14,37,56,67]. However, the ΔH value on Z1200 was constant in the whole range of loadings allowed, indicating that this sorbent is energetically homogeneous [4,8,17,56]. All ACs and A100 displayed an intermediate behavior. All NaY, NaX, and A100 showed ΔH values that increased after a certain CO2 coverage, which is due to the lateral interactions between the adsorbed molecules [17]. NaX has ΔH higher than that of NaY by 13 kJ/mol. This would result from the lower SAR of that zeolite (Table 1), as in the case of CO2 adsorption on Na-SSZ-13 with different SARs shown by Lobo and coworkers [67]. A significant dependence of ΔH for CO2, on the cage and window sizes of zeolite, has been reported [67]; however, such an effect is not anticipated, because the framework of NaY and NaX is the same. When CO2 is adsorbed on A100, an increase in ΔH occurs from a much lower loading (ca. 2 mmol/g) compared to the two zeolites. This may be associated with a structural shrinkage of A100 due to the interaction between CO2 and μ2-OH sites [11,32,56], thereby facilitating such adsorbate–adsorbate interactions. Consequently, the estimated ΔH values for CO2 adsorption again show the same order of the surface energetic heterogeneity of the adsorbents as that based on the shape of the measured isotherms.
Not only did the adsorbents used give significantly different selectivities of CO2 to CO, N2, and H2, but the selectivity factors also varied with the equilibrium pressures allowed (see Figure 7 and Table 2). The estimated selectivity factor is CO2/H2 > CO2/N2 > CO2/CO, and this order is maintained irrespective of the porous sorbent and chosen pressure, although the absolute values of each selectivity depend on the adsorbates and adsorbents. Such an order can be readily correlated to the difference (∆qm) in qm between CO2 and other gases, which means that the ∆qm value is the largest in a combination of CO2–H2 and the smallest in CO2–CO. Of course, the difference (∆α) in α between CO2 and the remaining adsorbates is apparently the same as the conclusion reached from the ∆qm values, but this ∆α effect is minor [4,5,68]. This clearly implies that the larger the ∆qm value, the greater the selectivity factor. Unfortunately, among the adsorbates used, a pair consisting of CO2 and CO—which are of particular interest in the efficient separation of CO2 from a gas mixture containing high concentrations of CO (near 38%)—as stated previously, gives the smallest ∆qm value, unlike the CO2–CH4 systems (∆qm = qm for CO2 because qm = 0 for CH4) that were the focus of many earlier studies [17,31,32,33].
Except for Z1200, all the adsorbents showed a decrease in the CO2/CO, CO2/N2, and CO2/H2 selectivity factors with equilibrated pressures, although the extent of such a decrease depended strongly on the sorbents (Figure 7 and Table 2), indicating energetically heterogeneous surfaces [4,69,70]. The Z1200, for which the selectivity factors increased at loadings below 2.5 bar and then remained constant, is characterized by energetically homogeneous surfaces [4,69,70]. This exceptional case is consistent not only with the isotherms in Figure 6 that are a linear function of pressure even for CO2 with the greatest α and qm values but also with the constant ΔH value at whole loadings of CO2 (Figure 11), as discussed earlier. Even the sorbents with energetically heterogeneous surfaces showed a significant difference in the extent of heterogeneity. For example, changes in the CO2/H2 selectivity factor, which would readily provide us with such a difference because of the largest ∆qm between the two molecules, with respect to equilibrated pressure, appeared in the order: NaY ≈ NaX >> A100 > GCN > MSP (Figure 7c). This shows that the surface energetic heterogeneity is the highest in the zeolites but the weakest in the ACs. However, the steep decrease in ΔH for CO2 gave the surface nonuniformity in the order: NaX > NaY ≈ GCN > MSP > A100 (Figure 11). Such a difference between the CO2/H2 selectivity factor- and ΔH-based surface heterogeneity orders may be due to higher H2 adsorption on ACs than on A100 (Figure 3, Figure 4 and Figure 5).
Of particular concern to us was the CO2/CO selectivity of the adsorbents used, because of the need for a PSA sorbent to preferentially separate CO2 from both CO2- and CO-rich gas streams in which such a selectivity may be critical. At 1 bar, the CO2/CO selectivity of A100 was 8.2, which is higher than those of the other adsorbents (Table 2). However, it is necessary to mention again that the suitability of porous solids as PSA adsorbents should be considered based on selectivity data at higher pressures but not exceeding 10 bar, such as 4–6 bar [14,20]. Z1200 has a selectivity factor of 6.2 at 6 bar, which is higher than that (2.6–4.3) shown for the conventional benchmarks, such as zeolites and ACs, and another MOF, i.e., A100 (Table 2). Despite this, MOFs have some limitations to their widespread use as adsorbents in PSA systems for CO2 separation from gas mixtures containing other light gases [63]. Representatively, MOF analogues are much less cost-effective because they are relatively expensive [11,63] and, to the best of our knowledge, have not yet been commercially proven in industrial PSA processes. Therefore, we think that NaY and MSP, which allow somewhat greater CO2/CO selectivity among the benchmarks, may be a good option. The common pitfall of ACs is a relatively low CO2/N2 selectivity [71], also indicated in our study, where MSP yields a CO2/N2 selectivity of 4.5 at 6 bar, which is lower than that for the NaY. However, such a disadvantage may be resolved in the case of a CO2-rich mixture stream, such as one with 38% CO2/4% N2, with both CO and H2 making up the remainder of the stream.

4.4. Working Capacity of CO2 Adsorption on the Adsorbents

One of the other parameters that should be assessed in adsorptive gas separation studies is the working capacity of adsorption of a target gas since the effectiveness of any microporous solid substance as a PSA sorbent is critically dependent on the extent of reloadings of the adsorbate in subsequent PSA cycles. We found that the estimated CO2 working capacity at 25 and 35 °C greatly varies with the adsorbents used (see Table 2 and Table S2), which is associated with their surface energetic heterogeneity and the population of high-affinity sites for CO2. MSP had the highest working capacity for CO2 adsorption at 25 °C, ca. 6.50 mmol/g. This value is much higher than the 2.2 mmol CO2/g working capacity reported for an LTA framework zeolite Rho showing the highest CO2/CH4 selectivity among zeolites studied to date [14] and, furthermore, is somewhat larger than or comparable to that (ca. 4–6.5 mmol/g) calculated at the chosen working pressure (1 and 6 bar) using CO2 isotherms reported for MOF-200 and MOF-210, MIL-101, PCN-68, and MOF-177 [28,29,30,34]. Even at 35 °C, MSP had a high CO2 adsorption working capacity (5.97 mmol/g), the largest among the six sorbents listed in Table S2. Thus, this AC is probably the most promising adsorbent because the working capacity determines the amount of CO2 processed in every adsorption–desorption cycle in PSA systems, although the selectivity factors are comparable to or less than those of the zeolites, A100, and Z1200, at 6 bar. In addition, the working capacity for CO2 adsorption on the MOFs at 25 and 35 °C was greater than that on NaY and NaX but was similar to that obtained for GCN (Table 2 and Table S2). Consequently, it is thought that among the porous sorbents, MSP may be a better choice for our later application in CO2 separation from biomass-gasification-based syngases even at the small sacrifice of CO2/N2 and CO2/H2 selectivities.

4.5. Regenerability of Adsorbents

We studied whether the successful regeneration of porous solid adsorbents should be an essential consideration for industrial PSA applications. We also studied CO2 and CO adsorption–desorption cycles with the zeolites and ACs using appropriate regeneration techniques, such as evacuation and fast pressure swing. Despite some earlier studies that have claimed difficulty in regenerating zeolite sorbents after CO2 adsorption under common PSA operating conditions [7,11,37], we could not see such an irreversibility in CO2 adsorption at 25 °C on faujasite-framework zeolites (NaY). CO adsorption on this material is also a fully reversible process (Figure 8a and Figure 10a).
The regenerability of the two ACs used is distinctive between themselves as well as between CO2 and CO adsorption. MSP showed fully reversible CO2 and CO adsorption regardless of the regeneration technique, whereas on GCN, such adsorption was possible for only CO (Figure 8b,c and Figure 10b,c). The irreversible behavior in the repeated CO2 adsorption on GCN, which occurred at pressure > 2.5 bar, may be due to the inorganic elements indigenous to the carbon source, particularly K, whose XRF intensity is much greater than that of peaks not only of the other metal impurities in GCN but also of the K existing in MSP (Figure 9). Commercial ACs include substantial amounts of mineral substances as impurity metals, such as Na, K, Si, Al, Fe, and Ca, depending on the original carbon-rich sources, typically wood, coal, lignin, coconut shell, and chemical resins, and the activation processes [19,72]. Even though such indigenous inorganic matters can be remarkably reduced using appropriate acid washing treatments with usually hot HCl and HF solutions, it is difficult to completely remove them, especially from micropores in ACs. Only a small amount of K (ca. 123 ppm) with traces of Fe, Ni, and Cr (all less than 25 ppm) was reported to exist in another MSP sample [73]. However, such metal impurity levels seem to have no appreciable effect on even CO2 adsorption, as observed in our measurements. It is proposed that inorganic metal impurities in carbonaceous sorbents should be identified and that their levels should be determined, if necessary, because of the significant dependence of such impurities on the original carbon sources and factory manufacturing processes.

5. Conclusions

Equilibrium adsorption of CO2, CO, N2, and H2 at 25 °C (including 35 °C for CO2) on zeolites, ACs, and MOFs was strongly associated not only with their physicochemical and surface features, but also with their molecular properties. All the adsorbents were microporous, and their Vmc, SBET, and surface polarity could influence the adsorption of the adsorbates but not directly correlate to them. For a given sorbent, the qm and α properties of the adsorbates played a critical role in determining their adsorption isotherm shapes, uptake, and selectivity factors, while with a chosen adsorbate, this adsorption strongly depended on the surface energetic heterogeneity of the sorbents. All the adsorbents, except for Z1200, were energetically heterogeneous but had a noticeable difference in terms of surface nonuniformity. The ΔH values calculated for CO2 adsorption on the six adsorbents and the changes with its coverage were in excellent agreement with the surface heterogeneity. Zeolites, NaY, and NaX exhibited relatively high CO2 adsorption at low pressures (near 1 bar), which was mainly because of strong extra-framework charge–quadrupole interactions between the Na cations and CO2. As adsorption pressure increases, the SBET of the adsorbents rather than their surface energetic heterogeneity may be more critical, as MSP shows the highest CO2 uptake at 6 bar. Neither of the adsorbents studied had a problem in regeneration, even when using fast pressure swing desorption, except for GCN, on which CO2 was irreversibly adsorbed at high pressures due to a significant quantity of metal impurities. The highest working capacity for CO2 adsorption (6.50 and 5.97 mmol/g at 25 and 35 °C, respectively) was obtained for MSP. This adsorbent may be a good candidate for separating CO2 from biomass-waste-driven syngases.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/su151511574/s1: Figure S1: Schematic of an adsorption system. Figure S2: N2 sorption isotherms of (a) NaY, (b) NaX, (c) GCN, (d) MSP, (e) A100, and (f) Z1200. Figure S3: Toth model fits to CO2 adsorption on the six adsorbents at 25, 35, and 40 °C. Table S1: Toth model parameters for CO2 adsorption isotherms on zeolites, ACs, and MOFs used. Table S2: Uptake of CO2 on microporous adsorbents at 25 and 35 °C and 6 bar and their working capacity.

Author Contributions

Conceptualization, M.H.K.; methodology, K.H.K. and M.H.K.; validation, K.H.K. and M.H.K.; visualization, K.H.K.; investigation, K.H.K. and M.H.K.; resources, K.H.K. and M.H.K.; data curation, K.H.K. and M.H.K.; writing—draft, K.H.K.; writing—review, editing, and modification, M.H.K.; supervision, M.H.K.; project administration, M.H.K.; funding acquisition, M.H.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was partly funded by the Daegu University Academic Research Program (grant #2022-0002).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data reported here are available upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Campoy, M.; Gómez-Barea, A.; Ollero, P.; Nilsson, S. Gasification of wastes in a pilot fluidized bed gasifier. Fuel Process. Technol. 2014, 121, 63–69. [Google Scholar] [CrossRef]
  2. Hanaoka, T.; Hiasa, S.; Edashige, Y. Syngas production by CO2/O2 gasification of aquatic biomass. Fuel Process. Technol. 2013, 116, 9–15. [Google Scholar] [CrossRef]
  3. Ammendola, P.; Raganati, F.; Chirone, R. Effect of operating conditions on the CO2 recovery from a fine activated carbon by means of TSA in a fluidized bed assisted by acoustic fields. Fuel Process. Technol. 2015, 134, 494–501. [Google Scholar] [CrossRef]
  4. Yang, R.T. Adsorbents: Fundamental and Applications; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2003. [Google Scholar]
  5. Kulprathipanja, S. Aspects of mechanisms, processes, and requirements for zeolite separation. In Zeolites in Industrial Separation and Catalysis; Kulprathipanja, S., Ed.; Wiley-VCH Verlag: Weinheim, Germany, 2010; pp. 203–228. [Google Scholar]
  6. Buckingham, J.; Reina, T.R.; Duyar, M.S. Recent advances in carbon dioxide capture for process intensification. Carbon Capture Sci. Technol. 2022, 2, 100031. [Google Scholar] [CrossRef]
  7. Sircar, S.; Golden, T.C. Purification of hydrogen by pressure swing adsorption. Sep. Sci. Technol. 2000, 35, 667–687. [Google Scholar] [CrossRef]
  8. Reddy, M.S.B.; Ponnamma, D.; Sadasivuni, K.K.; Kumar, B.; Abdullah, A.M. Carbon dioxide adsorption based on porous materials. RSC Adv. 2021, 11, 12658–12681. [Google Scholar] [CrossRef] [PubMed]
  9. Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.; Ferey, G. A rationale for the large breathing of the porous aluminum terephthalate (MIL-53) upon hydration. Chem. Eur. J. 2004, 10, 1373–1382. [Google Scholar] [CrossRef]
  10. Park, K.S.; Ni, Z.; Cote, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional chemical and thermal stability of zeolitic imidazolate frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef]
  11. Liu, J.; Thallapally, P.K.; McGrail, B.P.; Brown, D.R.; Liu, J. Progress in adsorption-based CO2 capture by metal-organic frameworks. Chem. Soc. Rev. 2012, 41, 2308–2322. [Google Scholar] [CrossRef]
  12. Li, J.R.; Kuppler, R.J.; Zhou, H.C. Selective gas adsorption and separation in metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1477–1504. [Google Scholar] [CrossRef]
  13. Bae, Y.S.; Snurr, R.Q. Development and evaluation of porous materials for carbon dioxide separation and capture. Angew. Chem. Int. Ed. 2011, 50, 11586–11596. [Google Scholar] [CrossRef]
  14. Palomino, M.; Corma, A.; Jorda, J.L.; Rey, F.; Valencia, S. Zeolite Rho: A highly selective adsorbent for CO2/CH4 separation induced by a structural phase modification. Chem. Commun. 2012, 48, 215–217. [Google Scholar] [CrossRef]
  15. Choi, S.; Drese, J.H.; Jones, C.W. Adsorbent materials for carbon dioxide capture from large anthropogenic point sources. ChemSusChem 2009, 2, 796–854. [Google Scholar] [CrossRef]
  16. Grande, C.A.; Lopes, F.V.S.; Ribeiro, A.M.; Loureiro, J.M.; Rodrigues, A.E. Adsorption of off-gases from steam methane reforming (H2, CO2, CH4, CO and N2) on activated carbon. Sep. Sci. Technol. 2008, 43, 1338–1364. [Google Scholar] [CrossRef]
  17. Sircar, S. Basic research needs for design of adsorptive gas separation processes. Ind. Eng. Chem. Res. 2006, 45, 5435–5448. [Google Scholar] [CrossRef]
  18. Mason, J.A.; Sumida, K.; Herm, Z.R.; Krishna, R.; Long, J.R. Evaluating metal-organic frameworks for post-combustion carbon dioxide capture via temperature swing adsorption. Energy Environ. Sci. 2011, 4, 3030–3040. [Google Scholar] [CrossRef]
  19. Jankowska, H.; Swiatkowski, A.; Choma, J. Active Carbon, Ellis Horwood and Wydawnictwa Naukowo-Techniczne; Ellis Horwood: Warsaw, Poland, 1991. [Google Scholar]
  20. Tagliabue, M.; Farrusseng, D.; Valencia, S.; Aguado, S.; Ravon, U.; Rizzo, C.; Corma, A.; Mirodatos, C. Natural gas treating by selective adsorption: Material science and chemical engineering interplay. Chem. Eng. J. 2009, 155, 553–566. [Google Scholar] [CrossRef]
  21. Yaghi, O.M.; Li, G.; Li, H. Selective binding and removal of guests in a microporous metal-organic framework. Nature 1995, 378, 703–706. [Google Scholar] [CrossRef]
  22. Ferey, G. Hybrid porous solids: Past, present, future. Chem. Soc. Rev. 2008, 37, 191–214. [Google Scholar] [CrossRef]
  23. Kitagawa, S.; Kitaura, R.; Noro, S. Functional porous coordination polymers. Angew. Chem. Int. Ed. 2004, 43, 2334–2375. [Google Scholar] [CrossRef]
  24. Rowsell, J.L.C.; Yaghi, O.M. Metal-organic frameworks: A new class of porous materials. Micropor. Mesopor. Mater. 2004, 73, 3–14. [Google Scholar] [CrossRef]
  25. Li, J.R.; Ma, Y.; McCarthy, M.C.; Sculley, J.; Yu, J.; Jeong, H.K.; Balbuena, P.B.; Zhou, H.C. Carbon dioxide capture-related gas adsorption and separation in metal-organic frameworks. Coord. Chem. Rev. 2011, 255, 1791–1823. [Google Scholar] [CrossRef]
  26. Ferey, G.; Mellot-Draznieks, C.; Serre, C.; Millange, F.; Dutour, J.; Surble, S.; Margiolaki, I. A chromium terephthalate-based solid with unusually large pore volumes and surface area. Science 2005, 309, 2040–2042. [Google Scholar] [CrossRef]
  27. Chae, H.K.; Siberio-Perez, D.Y.; Kim, J.; Go, Y.B.; Eddaoudi, M.; Matzger, A.J.; O’Keeffe, M.; Yaghi, O.M. A route to high surface area, porosity and inclusion of large molecules in crystals. Nature 2004, 427, 523–527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Yuan, D.; Zhao, D.; Sun, D.; Zhou, H.C. An isoreticular series of metal-organic frameworks with dendritic hexacarboxylate ligands and exceptionally high gas-uptake capacity. Angew. Chem. Int. Ed. 2010, 49, 5357–5361. [Google Scholar] [CrossRef]
  29. Furukawa, H.; Ko, N.; Go, Y.B.; Aratani, N.; Choi, S.B.; Choi, E.; Yazaydin, A.O.; Snurr, R.Q.; O’Keeffe, M.; Kim, J.; et al. Ultrahigh porosity in metal-organic frameworks. Science 2010, 329, 424–428. [Google Scholar] [CrossRef] [Green Version]
  30. Millward, A.R.; Yaghi, O.M. Metal-organic frameworks with exceptionally high capacity for storage of carbon dioxide at room temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef]
  31. Bastin, L.; Barcia, P.S.; Hurtado, E.J.; Silva, J.A.C.; Rodrigues, A.E.; Chen, B. A microporous metal organic framework for separation of CO2/N2 and CO2/CH4 by fixed-bed adsorption. J. Phys. Chem. C 2008, 112, 1575–1581. [Google Scholar] [CrossRef]
  32. Bourrelly, S.; Llewellyn, P.L.; Serre, C.; Millange, F.; Loiseau, T.; Ferey, G. Different adsorption behaviors of methane and carbon dioxide in the isotypic nanoporous metal terephthalates MIL-53 and MIL-47. J. Am. Chem. Soc. 2005, 127, 13519–13521. [Google Scholar] [CrossRef]
  33. Mu, B.; Li, F.; Walton, K.S. A novel metal-organic coordination polymer for selective adsorption of CO2 over CH4. Chem. Commun. 2009, 2493–2495. [Google Scholar] [CrossRef]
  34. Llewellyn, P.L.; Bourrelly, S.; Serre, C.; Vimont, A.; Daturi, M.; Hamon, L.; De Weireld, G.; Chang, J.S.; Hong, D.Y.; Hwang, Y.K.; et al. High uptakes of CO2 and CH4 in mesoporous metal-organic frameworks MIL-100 and MIL-101. Langmuir 2008, 24, 7245–7250. [Google Scholar] [CrossRef]
  35. Garces, S.I.; Villarroel-Rocha, J.; Sapag, K.; Korili, S.A.; Gil, A. Comparative study of the adsorption equilibrium of CO2 on microporous commercial materials at low pressures. Ind. Eng. Chem. Res. 2013, 52, 6785–6793. [Google Scholar] [CrossRef]
  36. McEwen, J.; Hayman, J.D.; Yazaydin, A.O. A comparative study of CO2, CH4 and N2 adsorption in ZIF-8, zeolite-13X and BPL activated carbon. Chem. Phys. 2013, 412, 72–76. [Google Scholar] [CrossRef]
  37. Siriwardane, R.V.; Shen, M.S.; Fisher, E.P.; Poston, J.A. Adsorption of CO2 on molecular sieves and activated carbon. Energy Fuels 2001, 15, 279–284. [Google Scholar] [CrossRef]
  38. Lopes, F.V.S.; Grande, C.A.; Ribeiro, A.M.; Loureiro, J.M.; Evaggelos, O.; Nikolakis, V.; Rodrigues, A.E. Adsorption of H2, CO2, CH4, CO, N2 and H2O in activated carbon and zeolite for hydrogen production. Sep. Sci. Technol. 2009, 44, 1045–1073. [Google Scholar] [CrossRef]
  39. Liang, Z.; Marshall, M.; Chaffee, A.L. CO2 adsorption-based separation by metal organic framework (Cu-BTC) versus zeolite (13X). Energy Fuels 2009, 23, 2785–2789. [Google Scholar] [CrossRef]
  40. Saha, D.; Bao, Z.; Jia, F.; Deng, S. Adsorption of CO2, CH4, N2O, and N2 on MOF-5, MOF-177, and zeolite 5A. Environ. Sci. Technol. 2010, 44, 1820–1826. [Google Scholar] [CrossRef]
  41. Krishna, R.; van Baten, J.M. A comparison of the CO2 capture characteristics of zeolites and metal-organic frameworks. Sep. Purif. Technol. 2012, 87, 120–126. [Google Scholar] [CrossRef]
  42. Saito, A.; Foley, H.C. Curvature and parametric sensitivity in models for adsorption in micropores. AIChE J. 1991, 37, 429–436. [Google Scholar] [CrossRef]
  43. Horvath, G.; Kawazoe, K. Method for the calculation of effective pore size distribution in molecular sieve carbon. J. Chem. Eng. Jap. 1983, 16, 470–475. [Google Scholar] [CrossRef] [Green Version]
  44. Kim, M.H.; Ebner, J.R.; Friedman, R.M.; Vannice, M.A. Dissociative N2O adsorption on supported Pt. J. Catal. 2001, 204, 348–357. [Google Scholar] [CrossRef]
  45. Kim, M.H.; Ebner, J.R.; Friedman, R.M.; Vannice, M.A. Determination of metal dispersion and surface composition in supported Cu–Pt catalysts. J. Catal. 2002, 208, 381–392. [Google Scholar] [CrossRef]
  46. Yang, W.H.; Kim, M.H. Catalytic reduction of N2O by H2 over well-characterized Pt surfaces. Korean J. Chem. Eng. 2006, 23, 908–918. [Google Scholar] [CrossRef]
  47. Sing, K.S.W.; Everett, D.H.; Haul, R.A.W.; Moscou, L.; Pierotti, R.A.; Rouquerol, J.; Siemieniewska, T. Reporting physisorption data for gas/solid systems with special reference to the determination of surface area and porosity. Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  48. Rege, S.U.; Yang, R.T. Corrected Horvath-Kawazoe equations for pore-size distribution. AIChE J. 2000, 46, 734–750. [Google Scholar] [CrossRef] [Green Version]
  49. Maurin, G.; Llewellyn, P.; Poyet, T.; Kuchta, B. Influence of extra-framework cations on the adsorption properties of X-faujasite systems: Microcalorimetry and molecular simulations. J. Phys. Chem. B 2005, 109, 125–129. [Google Scholar] [CrossRef]
  50. Heymans, N.; Vaesen, S.; De Weireld, G. A complete procedure for acidic gas separation by adsorption on MIL-53 (Al). Micropor. Mesopor. Mater. 2012, 154, 93–99. [Google Scholar] [CrossRef]
  51. Xuan-Dong, D.; Vinh-Thang, H.; Kaliaguine, S. MIL-53 (Al) mesostructured metal-organic frameworks. Micropor. Mesopor. Mater. 2011, 141, 135–139. [Google Scholar]
  52. Deniz, E.; Karadas, F.; Patel, H.A.; Aparicio, S.; Yavuz, C.T.; Atilhan, M. A combined computational and experimental study of high pressure and supercritical CO2 adsorption on Basolite MOFs. Micropor. Mesopor. Mater. 2013, 175, 34–42. [Google Scholar] [CrossRef]
  53. Blanco-Brieva, G.; Campos-Martin, J.M.; Al-Zahrani, S.M.; Fierro, J.L.G. Effectiveness of metal-organic frameworks for removal of refractory organo-sulfur compound present in liquid fuels. Fuel 2011, 90, 190–197. [Google Scholar] [CrossRef]
  54. Ferey, G.; Serre, C. Large breathing effects in three-dimensional porous hybrid matter: Facts, analyses, rules and consequences. Chem. Soc. Rev. 2009, 38, 1380–1399. [Google Scholar] [CrossRef]
  55. Almasoudi, A.; Mokaya, R. Preparation and hydrogen storage capacity of templated and activated carbons nanocast from commercially available zeolitic imidazolate framework. J. Mater. Chem. 2012, 22, 146–152. [Google Scholar] [CrossRef]
  56. Ramsahye, N.A.; Maurin, G.; Bourrelly, S.; Llewellyn, P.; Loiseau, T.; Ferey, G. Charge distribution in metal organic framework materials: Transferability to a preliminary molecular simulation study of the CO2 adsorption in the MIL-53 (Al) system. Phys. Chem. Chem. Phys. 2007, 9, 1059–1063. [Google Scholar] [CrossRef]
  57. Ruthven, D.M. Fundamentals of adsorption equilibrium and kinetics in microporous solids. Mol. Sieves 2008, 7, 1–43. [Google Scholar]
  58. Toth, J. Uniform and thermodynamically consistent interpretation of adsorption isotherms. In Adsorption: Theory, Modeling, and Analysis; Toth, J., Ed.; Marcel Dekker, Inc.: New York, NY, USA, 2002; pp. 1–104. [Google Scholar]
  59. Dunne, J.A.; Rao, M.; Sircar, S.; Gorte, R.J.; Myers, A.L. Calorimetric heats of adsorption and adsorption isotherms: 2. O2, N2, Ar, CO2, CH4, C2H6, and SF6 on NaX, H-ZSM-5, and Na-ZSM-5 zeolites. Langmuir 1996, 12, 5896–5904. [Google Scholar] [CrossRef]
  60. Pirngruber, G.D.; Raybaud, P.; Belmabkhout, Y.; Cejka, J.; Zukal, A. The role of the extra-framework cations in the adsorption of CO2 on faujasite Y. Phys. Chem. Chem. Phys. 2010, 12, 13534–13546. [Google Scholar] [CrossRef]
  61. Perez-Pellitero, J.; Amrouche, H.; Siperstein, F.R.; Pirngruber, G.; Nieto-Draghi, C.; Chaplais, G.; Simon-Masseron, A.; Bazer-Bachi, D.; Peralta, D.; Bats, N. Adsorption of CO2, CH4, and N2 on zeolitic imidazolate frameworks: Experiments and simulations. Chem. Eur. J. 2010, 16, 1560–1571. [Google Scholar] [CrossRef]
  62. Couck, S.; Denayer, J.F.M.; Baron, G.V.; Remy, T.; Gascon, J.; Kapteijn, F. An amine-functionalized MIL-53 metal-organic framework with large separation power for CO2 and CH4. J. Am. Chem. Soc. 1999, 131, 6326–6327. [Google Scholar] [CrossRef]
  63. D’Alessandro, D.M.; Smit, B.; Long, J.R. Carbon dioxide capture: Prospects for new materials. Angew. Chem. Int. Ed. 2010, 49, 6058–6082. [Google Scholar] [CrossRef] [Green Version]
  64. Boddenberg, B.; Rakhmatkariev, G.U.; Wozniak, A.; Hufnagel, S. A calorimetric and statistical mechanics study of ammonia adsorption in zeolite NaY. Phys. Chem. Chem. Phys. 2004, 6, 2494–2501. [Google Scholar] [CrossRef]
  65. Ward, J.W.; Habgood, H.W. The infrared spectra of carbon dioxide adsorbed on zeolite X. J. Phys. Chem. 1966, 70, 1178–1182. [Google Scholar] [CrossRef]
  66. Ania, C.O.; Parra, J.B.; Menendez, J.A.; Pis, J.J. Effect of microwave and conventional regeneration on the microporous and mesoporous network and on the adsorptive capacity of activated carbons. Micropor. Mesopor. Mater. 2005, 85, 7–15. [Google Scholar] [CrossRef] [Green Version]
  67. Pham, T.D.; Liu, Q.; Lobo, R.F. Carbon dioxide and nitrogen adsorption on cation-exchanged SSZ-13 zeolites. Langmuir 2013, 29, 832–839. [Google Scholar] [CrossRef] [PubMed]
  68. Herm, Z.R.; Krishna, R.; Long, J.R. CO2/CH4, CH4/H2 and CO2/CH4/H2 separations at high pressures using Mg2(dobdc). Micropor. Mesopor. Mater. 2012, 151, 481–487. [Google Scholar] [CrossRef]
  69. Sircar, S.; Rao, M.B. Effect of adsorbate size on adsorption of gas mixtures on homogeneous adsorbents. AIChE J. 1999, 45, 2657–2661. [Google Scholar] [CrossRef]
  70. Sircar, S.; Myers, A.L. Gas separation by zeolites. In Handbook of Zeolite Science and Technology; Auerbach, S.M., Carrado, K.A., Dutta, P.K., Eds.; Marcel Dekker Inc.: New York, NY, USA, 2003; pp. 1354–1404. [Google Scholar]
  71. Radosz, M.; Hu, X.D.; Krutkramelis, K.; Shen, Y.Q. Flue-gas carbon capture on carbonaceous sorbents:  toward a low-cost multifunctional carbon filter for “green” energy producers. Ind. Eng. Chem. Res. 2008, 47, 3783–3794. [Google Scholar] [CrossRef]
  72. Wang, S.; Lu, G.Q. Effects of acidic treatments on the pore and surface properties of Ni catalyst supported on activated carbon. Carbon 1998, 36, 283–292. [Google Scholar] [CrossRef]
  73. Kim, J.A.; Park, I.S.; Seo, J.H. A development of high power activated carbon using the KOH activation of soft carbon series cokes. Trans. Electr. Electron. Mater. 2014, 15, 81–86. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Adsorption isotherms of CO2, CO, N2, and H2 on NaY at 25 °C. Solid lines are just to guide the eyes.
Figure 1. Adsorption isotherms of CO2, CO, N2, and H2 on NaY at 25 °C. Solid lines are just to guide the eyes.
Sustainability 15 11574 g001
Figure 2. Adsorption isotherms of CO2, CO, N2, and H2 on NaX at 25 °C. Solid lines are just to guide the eyes.
Figure 2. Adsorption isotherms of CO2, CO, N2, and H2 on NaX at 25 °C. Solid lines are just to guide the eyes.
Sustainability 15 11574 g002
Figure 3. Adsorption isotherms of CO2, CO, N2, and H2 on GCN at 25 °C. Solid lines are just to guide the eyes.
Figure 3. Adsorption isotherms of CO2, CO, N2, and H2 on GCN at 25 °C. Solid lines are just to guide the eyes.
Sustainability 15 11574 g003
Figure 4. Adsorption isotherms of CO2, CO, N2, and H2 on MSP at 25 °C. Solid lines are just to guide the eyes.
Figure 4. Adsorption isotherms of CO2, CO, N2, and H2 on MSP at 25 °C. Solid lines are just to guide the eyes.
Sustainability 15 11574 g004
Figure 5. Adsorption isotherms of CO2, CO, N2, and H2 on A100 at 25 °C. Solid lines are just to guide the eyes.
Figure 5. Adsorption isotherms of CO2, CO, N2, and H2 on A100 at 25 °C. Solid lines are just to guide the eyes.
Sustainability 15 11574 g005
Figure 6. Adsorption isotherms of CO2, CO, N2, and H2 on Z1200 at 25 °C. Solid lines are just to guide the eyes.
Figure 6. Adsorption isotherms of CO2, CO, N2, and H2 on Z1200 at 25 °C. Solid lines are just to guide the eyes.
Sustainability 15 11574 g006
Figure 7. Selectivity of CO2 to CO, N2, and H2 for microporous adsorbents with respect to pressure: (a) CO2/CO, (b) CO2/N2, and (c) CO2/H2.
Figure 7. Selectivity of CO2 to CO, N2, and H2 for microporous adsorbents with respect to pressure: (a) CO2/CO, (b) CO2/N2, and (c) CO2/H2.
Sustainability 15 11574 g007
Figure 8. Repeated measurements of CO2 adsorption at 25 °C and pressures of up to ca. 7.5 bar with selected sorbents: (a) NaY, (b) GCN, and (c) MSP. An open symbol in (a) indicates CO2 adsorption on a surface subjected to a pressure swing to 1 from ca. 7.5 bar after the third adsorption.
Figure 8. Repeated measurements of CO2 adsorption at 25 °C and pressures of up to ca. 7.5 bar with selected sorbents: (a) NaY, (b) GCN, and (c) MSP. An open symbol in (a) indicates CO2 adsorption on a surface subjected to a pressure swing to 1 from ca. 7.5 bar after the third adsorption.
Sustainability 15 11574 g008
Figure 9. Wavelength-dispersive XRF spectra for fresh samples of the activated carbon sorbents used: (a) GCN, and (b) MSP. The original intensity of a peak at 2θ = 136.55° is marked with an asterisk due to K, as an impurity metal, being reduced by 90% for a ready comparison with other peaks.
Figure 9. Wavelength-dispersive XRF spectra for fresh samples of the activated carbon sorbents used: (a) GCN, and (b) MSP. The original intensity of a peak at 2θ = 136.55° is marked with an asterisk due to K, as an impurity metal, being reduced by 90% for a ready comparison with other peaks.
Sustainability 15 11574 g009
Figure 10. Repeated measurements of CO adsorption at 25 °C and pressures of up to ca. 7.5 bar with selected sorbents: (a) NaY, (b) GCN, and (c) MSP. An open symbol in (a) indicates CO adsorption on a surface subjected to a pressure swing from ca. 7.3 bar to 1 bar after the third adsorption.
Figure 10. Repeated measurements of CO adsorption at 25 °C and pressures of up to ca. 7.5 bar with selected sorbents: (a) NaY, (b) GCN, and (c) MSP. An open symbol in (a) indicates CO adsorption on a surface subjected to a pressure swing from ca. 7.3 bar to 1 bar after the third adsorption.
Sustainability 15 11574 g010
Figure 11. Isosteric heats of CO2 adsorption on NaY, NaX, GCN, MSP, A100, and Z1200.
Figure 11. Isosteric heats of CO2 adsorption on NaY, NaX, GCN, MSP, A100, and Z1200.
Sustainability 15 11574 g011
Table 1. Textural properties of the zeolites, ACs, and MOFs used.
Table 1. Textural properties of the zeolites, ACs, and MOFs used.
AdsorbentCategorySAR aSBETdmc b (Å)Pore Volume (cm3/g)fmcc
(m2/g)VmcVmsVt
NaYZeolite2.38477.850.3130.0280.3340.94
NaX dZeolite1.37247.580.2720.0970.3570.76
GCNAC-11325.02 e0.4390.0760.4790.92
MSPAC-25085.53 e0.9910.2991.1460.86
A100 fMOF-83810.960.3610.8991.1090.33
Z1200 gMOF-130112.270.5530.1250.7390.75
AC: activated carbon; MOF: metal–organic framework; SAR: Si-to-Al ratio; SBET: BET surface area; dmc: micropore diameter; fmc: micropore fraction; Vmc: micropore volume; Vms: mesopore volume; Vt: total pore volume; “-”: not applicable. a Using inductively coupled plasma (ICP) measurements. b Using the Saito–Foley cylindrical pore model. c Defined as fmc = Vmc/Vt. d Commonly called “zeolite 13X”. e Using the original Horvath–Kawazoe slit pore model. f A commercial form isostructural to MIL-53(Al) in Ref. [9]. g A commercial form isostructural to ZIF-8 in Ref. [10].
Table 2. Uptake of CO2, CO, N2, and H2 on microporous adsorbents at 25 °C and 6 bar and their selectivity factors and working capacity for CO2 adsorption.
Table 2. Uptake of CO2, CO, N2, and H2 on microporous adsorbents at 25 °C and 6 bar and their selectivity factors and working capacity for CO2 adsorption.
AdsorbentGas Uptake (mmol/g) a Selectivity Factor a,bWorking
Capacity (mmol/g) c
CO2CON2H2 CO2/COCO2/N2CO2/H2
NaY6.70 (5.60)1.95 (1.10)1.08 (0.23)0.13 (0.02) 3.4 (5.1)6.2 (24.4)51.5 (280)1.10
NaX5.38 (4.60)2.08 (1.12)1.43 (0.65)0.12 (0.02) 2.6 (4.1)3.8 (4.1)44.8 (230)0.78
GCN5.87 (2.75)1.94 (0.57)1.48 (0.40)0.19 (0.04) 3.0 (4.8)4.0 (6.9)30.9 (68.7)3.12
MSP9.75 (3.25)2.72 (0.68)2.19 (0.50)0.31 (0.06) 3.6 (4.8)4.5 (6.5)31.4 (54.1)6.50
A1005.18 (2.20)1.22 (0.27)0.89 (0.19)0.11 (0.02) 4.3 (8.2)5.8 (11.6)47.1 (110)2.98
Z12003.95 (0.66)0.64 (0.12)0.52 (0.09)0.10 (0.02) 6.2 (5.5)7.6 (7.3)39.5 (33)3.29
a Values in parenthesis are at 1 bar. b Based on each equilibrium uptake on the corresponding pure-gas isotherms. c Difference between CO2 uptake values at 1 and 6 bar.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kim, K.H.; Kim, M.H. Adsorption of CO2, CO, H2, and N2 on Zeolites, Activated Carbons, and Metal-Organic Frameworks with Different Surface Nonuniformities. Sustainability 2023, 15, 11574. https://doi.org/10.3390/su151511574

AMA Style

Kim KH, Kim MH. Adsorption of CO2, CO, H2, and N2 on Zeolites, Activated Carbons, and Metal-Organic Frameworks with Different Surface Nonuniformities. Sustainability. 2023; 15(15):11574. https://doi.org/10.3390/su151511574

Chicago/Turabian Style

Kim, Kang Hun, and Moon Hyeon Kim. 2023. "Adsorption of CO2, CO, H2, and N2 on Zeolites, Activated Carbons, and Metal-Organic Frameworks with Different Surface Nonuniformities" Sustainability 15, no. 15: 11574. https://doi.org/10.3390/su151511574

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop