Next Article in Journal
SFC and CE—A Comparison of Two Orthogonal Methods for the Analysis of Dihydrochalcones in Apple Leaves
Previous Article in Journal
Green-Engineered Barrier Creams with Montmorillonite-Chlorophyll Clays as Adsorbents for Benzene, Toluene, and Xylene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

CO2 Adsorption on Natural Zeolites from Puebla, México, by Inverse Gas Chromatography

by
Miguel Angel Hernandez
1,*,
Gabriela Itzel Hernandez
2,
Roberto Portillo
3,
Efraín Rubio
4,
Vitalii Petranovskii
5,
Karin Montserrat Alvarez
1,
Ma de los Angeles Velasco
1,
Juana Deisy Santamaría
6,
Mario Tornero
1 and
Laura Alicia Paniagua
7
1
Departament of Zeolites Research, Postgraduate in Agroecology, ICUAP, Meritorious Autonomous University of Puebla, Puebla City 72570, Mexico
2
Department of Process Engineering, Metropolitan Autonomous University-Iztapalapa, Mexico City 09340, Mexico
3
Faculty of Chemical Sciences, Meritorious Autonomous University of Puebla, Puebla City 72570, Mexico
4
University Center for Linking and Technology Transfer, Meritorious Autonomous University of Puebla, Puebla City 72570, Mexico
5
Nanosciences and Nanotechnology Center, UNAM, Ensenada 22860, Mexico
6
Faculty of Chemical Engineering, Meritorious Autonomous University of Puebla, Puebla City 72570, Mexico
7
Faculty of Electronic Sciences, Meritorious Autonomous University of Puebla, Puebla City 72570, Mexico
*
Author to whom correspondence should be addressed.
Separations 2023, 10(4), 238; https://doi.org/10.3390/separations10040238
Submission received: 23 December 2022 / Revised: 13 February 2023 / Accepted: 22 February 2023 / Published: 4 April 2023

Abstract

:
The applicability of clinoptilolite zeolites in controlling the emission of greenhouse gases (GHGs) such as CO2, the most significant GHG, is investigated herein. In this research, Mexican natural zeolites (ATN) originating from an Atzinco deposit in the state of Puebla were used. Samples of modified clinoptilolite (ATH4, ATH3, ATH2 and ATH1) were obtained from the starting material by acid treatment of various intensities. Inverse gas chromatography was used to evaluate CO2 adsorption in clinoptilolite, natural and chemically modified. Adsorption of CO2 was investigated in the temperature range of 433–573 K, using a TCD detector, and He as a carrier gas. The experimental CO2 adsorption data were processed by Freundlich and Langmuir equations. The degree of interaction between CO2 and the dealuminated clinoptilolite samples was examined through the evaluation of the isosteric enthalpy of adsorption. This calculation was made by using the Clausius–Clapeyron equation, which established the following sequence: ATH1 > ATH2 > ATH4 > ATN > ATH3. The nanoporosity of these clinoptolite zeolites from new deposit in sedimentary rocks was studied through HRADS adsorption of N2. Simultaneously, these zeolites were, respectively, characterized by XRD, EDS, and SEM. Micropores are described by the Dubinin–Asthakov distribution. Various adsorption mechanisms that occur in these nanoporous materials at different relative pressures can be visualized. The quantitative determination of starting mineral is described as: Ca-Clinoptilolite (88.76%) >> Montmorillonite (11.11%) >> quartz (0.13%). The Si/Al molar ratio after acid treatment is: ATH4 > ATH2 > ATN > ATH3 > ATH1. The Langmuir specific surface area (ASL) varies as follows: ATN > ATH2 > ATH4 > ATH3 > ATH1. At the same time, the VΣ values are as follows: ATN > ATH4 > ATH3 > ATH1 > ATH2.

1. Introduction

One of the biggest environmental challenges nowadays refers to the presence of carbon dioxide (CO2). Currently, the increase of this gas is a significant concern due to its gradual accumulation in the atmospheric concentration. This increment is the result of fossil fuels high consumption. This utilization is related to an increase in the concentration of CO2 in the atmosphere and the global climate change. Due to this problem, it is necessary to significantly counteract CO2 concentrations from the current levels [1]. There are several cutting-edge technologies used to reduce CO2 concentrations. Some techniques are focused on stabilizing the CO2 concentration in the atmosphere. For example, amine solutions for absorption, membrane separation and adsorption-based separation. These methods controlled the emission of CO2 from various sources. There exist different materials to selectively adsorb this gas [2,3]. Among these materials, there are hydrotalcites, inorganic oxides, activated carbons, MCM, MOFs, TIFs and synthetic (NaX, FAU and LTA) and natural (ERI, CHA and HEU) zeolites [2,3]. To date, the clinoptilolite zeolite is one of the most promising zeolites to carry out more selective adsorption processes with respect to CO2. The structural code for this zeolite is HEU, as per the International Zeolite Association, IZA [4]. This zeolite possesses remarkable adsorption characteristics, for instance, its accessibility, and purity. Additionally, it has the possibility to increase their adsorption capacity. This enhancement can be made by integrating certain cations through ion exchange, and/or via nanodeposits creation in the internal arrangement of this zeolite.
The substitution by ion exchange of Na+ ions in natural clinoptilolite by either Ca2+ or H+ ions improve the capacity (i.e., the accessible pore volume). Chemical treatments, via ion exchange and dealumination of zeolite result in the development of mesopores, known as secondary porosity. Clinoptilolite is a hydrophilic zeolite which structure is based on a three-dimensional porous system form by three types of channels [5]. The first two channels are two parallel channels along [001] axis: channel A with 10-member ring (0.31 × 0.75 nm) in cross-section, and channel B with 8-MR ring (0. 36 × 0.46 nm}. These A and B channels are perpendicularly intersected along [100] axis by channel C. The features of channel C are an 8-MR ring (0.28 × 0.47 nm). The channel sizes are variable due to the considerable flexibility of the framework), see Figure 1.
Natural zeolites are elaborated adsorption structures. They are constituted from different crystalline phases. This composition is related to the materials that constitute this type of mineral. The presence of micropores (pore diameter, W < 2 nm) is related to the sole presence of the zeolite. Meanwhile, the mesopores (W = 2–50 nm), and macropores (W > 50 mm) [6,7] are associated with the presence of accompanying materials. These additional materials are usually clays and quartz. During an adsorption process in this type of material, a micropore filling is carried out at very low relative pressures [7,8]. This mechanism involves a primary filling and a secondary filling [7,8]. As the relative pressure increases, the formation of the monolayer occurs. Secondly, there is a gradual increase in relative pressure and larger pores (mesopores) are occupied. The phenomenon of capillary condensation takes place in the macropores. This event occurs when the equilibrium pressure values tend to the saturation pressure values. Consequently, the formation of the adsorbate meniscus appears.
Gas–solid chromatography (GSC) is one of the suggested methods to study microporous materials with micropores at low coverage degrees. This method is called inverse gas chromatography (IGC) when it is applied to study solid surface properties. IGC has been used extensively to study catalytic reactions and textural parameters (such as surface area and porosity). Additionally, IGC has been broadly used to analyze various types of materials [9]. IGC can be experimentally configured for finite or infinite dilution concentrations of adsorbate [10]. According to these last concepts of the IGC method, in the case of infinite dilution concentrations, a few molecules are injected into the column to approach zero surface coverage. Considering this circumstance, the lateral interactions between the adsorbed molecules on the surface can be ignored. Hence, the thermodynamic functions only change with the adsorbate–adsorbent interactions. Adsorption can be considered to occur in the linear zone of the adsorption isotherm in the Henry’s law area [11]. The attainment of the Henry’s law region is indicated by certain characteristics. Among these features, there are the symmetry of the chromatographic peaks and the constancy of the retention times. These times, measured over a significant range of the sample sizes. Therefore, the net retention volumes for a specific adsorbate are non-dependent of its gas phase concentration [12].
This study examined the adsorption capacity of natural clinoptilolite from a new deposit in Mexico. Furthermore, it also examines the modifications by decationation. Finally, it also analyzes the possibly concomitant dealumination during acid treatment. The original material and the resulting modified samples were studied using XRD, EDS, and SEM. To assess their texture properties, HRADS measurements were carried out. By using the Dubinin–Astakhov (D-A) model with the HRADS results, the micropores principal parameters (W0 and E) were estimated.
The adsorption of CO2 in a series of natural and decationated zeolites was studied. The adsorptions were analyzed using the Langmuir and Freundlich equations. Meanwhile, the degree of interaction between this CO2 gas and samples of natural and decationated clinoptilolite were studied through the evaluation of the isosteric enthalpies of adsorption. Those enthalpies were analyzed with the Clausius–Clapeyron equation at zero degrees of coverage by IGC method.

2. Results

2.1. XRD

In Figure 2, the diffraction patterns of natural zeolites studied are shown. The phases observed for each zeolite sample are listed in Table 1. Quantitative determination of the percentage of these phases in ATN was conducted by using the High Score Plus 3.0e computer program. Additionally, appropriately processing diffraction patterns (XRD) and the current phases atomic coordinates were also calculated. Table 1 demonstrates the results of these estimates.
In Figure 3, the X-ray diffraction patterns are observed in the chemically modified zeolites (ATH1, ATH2, ATH3 and ATH4).

2.2. EDS

Table 2 shows the representative analyses of chemical composition of clinoptilolites studied in this research.
The chemical treatments to which the natural zeolite ATN are subjected allow us to establish the following sequences:
SiO2:
ATH1 > ATH3 > ATH2 > ATH4 > ATZN; Al2O3: ATH3 > ATH1 > ATH2 > ATH4 > ATZN;
F2O3:
ATH4 > ATH2 > ATH3 > ATH1 > ATZN; CaO: ATZN > ATH3 > ATH2 > ATH4 > ATH1;
MgO:
ATH2 > ATZN > ATH4 > ATH3 > ATH1; K2O: ATH3 > ATH1 > ATH2 > ATH4 > ATZN;
TiO2:
ATH4 = ATH1 > ATH2 > ATH3 < ATZN; FeO: ATH4 > ATH2 > ATZN > ATH3 > ATH1 and
Si/Al:
ATH4 > ATH2 > ATZN > ATH3 > ATH1.

2.3. SEM

Micrographs of clinoptilolite samples displayed classic crystalline morphology, Figure 4.

2.4. N2 Adsorption

Figure 5a,b shows the N2 adsorption isotherms of the studied zeolites. Table 3 reports the results of the textural studies of the analyzed samples.

2.4.1. Pore Size Distribution (PSD), DA and BJH Approaches

D-A Approach

The matching nanopores size distribution was obtained from the N2 adsorption–desorption isotherms at 77 K. The results of these calculations, Figure 6, demonstrate the uniformly distribution of the nanochannels, which diameters vary from 0.956 to 0.791 nm. This figure shows the intensity of these signals’ associates with the Al presence in the arrangement of these natural zeolites, (shown in Table 4, 6th column) [13].

BJH Equation

The results when applying the BJH method in the ATN-ATH zeolites are observed in Figure 7, in which it can be observed there are predominant signals in the area of pore diameters Dp = 2167 nm (ATH2), 3033 nm (ATH4) and 3606 nm (ATN). It is also observed that these distributions become less acute depending on the chemical treatment to which this type of zeolite is subjected. Again, this effect is attributed to the generation of these types of pores during dealumination treatments. ATH1 and AH3 zeolites were not subjected to BJH analysis.

2.5. Adsorption of CO2

The method of inverse gas chromatography (IGC) was used to estimate the adsorption isotherms based on the elution curves of adsorbate on the zeolites under study. These CO2 adsorption isotherms were measured in a region of low adsorbate concentrations; hence, adsorption on the mesopore surface can be practically neglected. Consequently, the thermodynamics parameters that characterize the microporous structure were calculated from the CO2 adsorption isotherms, without any correction for adsorption on the mesopore surface. CO2 adsorption isotherms at different temperatures on ATN and ATH zeolites are presented in parts a–e in Figure 8. Additionally, the degree of interaction of CO2 with zeolite was analyzed from the evolution of the isosteric enthalpy of adsorption -qst, Figure 9. The values of the Freundlich constants KF and n, Henry’s constants KH, and the Langmuir monolayer capacity are shown in Table 5. While the isosteric enthalpies of adsorption -qst are given in Table 6. From the measured retention volume through the column, different thermodynamic parameters can be deduced at the molecular level [14].
From Table 5, the following sequence is established for monolayer capacity (am, mmol g−1) at the studied temperatures: 537 K: ATH2 > ATH3 > ATH1 > ATN > ATH4; 543 K: ATH1 > ATH2 > ATH4 > ATH3 > ATHN; 513 K: ATH4 > ATH1 > ATH2 > ATH3 > ATHN and 433 K: ATN > ATH2 > ATH4 > ATH1 > ATH3. While those minor values are obtained in the ATN sample at 473 K, the highest values correspond to the same sample at 433 K. The Langmuir methodology applies moderately well for all except the ones regarding adsorption at 573 and 473 K on ATH4, and ATH2 and ATH4 at 433 K.
Table 6. Adsorption isosteric enthalpy (-qst, kJ mol−1) of CO2 on ATN and ATH clinoptilolites.
Table 6. Adsorption isosteric enthalpy (-qst, kJ mol−1) of CO2 on ATN and ATH clinoptilolites.
Sample-qst (kJ mol−1)
ERIN [15]18.8
ERINa226.2
Z17 [16]7.7
ZSLP [17]2.8
ATN *2.2
ATH1 *4.3
ATH2 *2.7
ATH3 *2
ATH4 *2.6
SNN100 [18]20
13X [19]40
MOF-5 [20]30
* This work.

3. Discussion

3.1. XRD

In Figure 2, it can be observed the natural zeolite (ATN) diffraction patterns under study. Table 1 presents the results of the estimates of the calculations of phases present. These samples contain clinoptilolite (JCPDS card 3 0427) [21] and, in smaller amounts, impurities (e.g., montmorillonite (JCPDS 29-1498), quartz (peak at ~ 27° 2θ, JCPDS 3-0427)). Figure 3 shows the X-ray diffraction patterns observed in the chemically modified zeolites (ATH1, ATH2, ATH3 and AH4). The distinguishing crystalline peaks of the clinoptilolite zeolite appear at the succeeding 2θ diffraction angles: 2θ values of 9.86°, 11.06°, 13.03°, 14.82°, 16.86°, 17.2°, 19.04°, 22.35°, 25.04°, 28.09°, 31.71°, 32.67°. It can be seen from this figure that peaks of clinoptilolite prevail. To clearly represent these signals in zeolite samples, the diffraction patterns were divided into 3 zones with a more detailed scale: zone (a) 2θ = 5–20°, (b) 2θ = 20–35°, (c) 2θ = 35–50° and, (d) in the zone of 2θ = 5–50°, the overall picture can be observed.

3.2. EDS

Representative analyses of the chemical composition of the clinoptilolites studied in this research are shown in Table 2. From these results, the parent zeolite has a Si/Al ratio > 5; such composition is typical for zeolites, which can be a result of settlings in saltwater bodies such as saline lakes and seas. It is known that amid clinoptilolite zeolites with a high Si content, that is, with a Si/Al ratio > 4, K-Clinoptilolite is the most widespread; this zeolite is dominant among the rocks produced in deep water zones. Nevertheless, Ca-Clinoptilolite and Na Clinoptilolite zeolites can be produced in a varied range of environmental factors. They can emerge in volcanic rocks, hydrothermal active systems, equally in cracks and fissures of volcanic rocks. Nonetheless, the Fe impurity is most likely Fe3+, which is in tetrahedral sites of crystalline framework, but its large amounts can also be associated with hematite impurities [22].

3.3. SEM

Figure 4 presents clinoptilolite crystals images at distinct magnifications. These images show that the crystals of this zeolite have different sizes and geometries. In this figure, it is possible to observe the crystals’ presence. These crystals have undefined, or irregular (anhedral) faces, which are typical of this zeolite [23,24]. Moreover, from Figure 4, it is observed the existence of a cluster of lamellar pores.

3.4. N2 Adsorption

Regarding the specific Langmuir surface area, estimated by the ASL method, the following sequence order is established: (ASL) is: ATN > ATH2 > ATH4 > ATH3 > ATH1. While the VΣ values are as follows: ATN > ATH4 > ATH3 > ATH1 > ATH2. A comparable performance is demonstrated by the total pore volume VΣ. Meanwhile, the textural properties values of the clinoptilolite zeolites are well described with the Langmuir equation.

D-A Approach

The same Table 4 shows the optimized values of W0, micropore volume, and the characteristic adsorption energy E0 values, obtained through the DA equation. The results show an increase in the amount of Al leads to a greater accessibility of the nanopores. The acquired values for E0 increase while Al2O3 decreases and are also affected by cations. The values found show the relationship between E0 and the Si/Al ratio, that is, its dependence on the Al2O3 content.
The isomorphous substitution of Si(IV) by Al(III) in the tetrahedral units of the zeolite crystalline structure results in the appearance of an unbalanced negative charge. Additionally, it also shows the presence of nonstructural, charge-balancing exchangeable cations located into the voids of the zeolite structure. It is evident that a change in the Si/Al ratio and the kind of counter-cations altered the adsorption potential. The concentration of Al atoms and the electrostatic interaction with cations affect the electric fields within the nanopores of clinoptilolite zeolite [25]. Therefore, they affect the interaction with sorbed molecules. From the size distribution of nanopores in Figure 7, it can be said the intensity of these signals is related with the presence of Al in these zeolites arrangement, (Table 5, 6th column).

3.5. CO2 Adsorption

In Figure 8a–e, three important points can be pointed: (i) the adsorbed quantities of CO2 gas on ATN and ATH zeolites are comparable, (ii) the adsorption isotherms at 573 K are better described experimentally and (iii) the adsorbed amount of CO2 decreases with temperature. Figure 9 exhibits the isosteric enthalpy of adsorption on ATN and ATH that rises with increasing surface-loading of CO2. For all zeolites, the adsorption isosteric enthalpy reduces with growing adsorbate loading. The initial acute reduction in -qst may be caused by the heterogeneity site. Nevertheless, a constant gradual decrease in -qst is expected at higher adsorbate loadings. This phenomenon caused by the repellent interactions happening among the adsorbed CO2 molecules. As a comparison, Table 6 includes values of isosteric enthalpy of adsorption of CO2 on erionite, clinoptilolite, 13X zeolites, SNN-100 hybrid materials and metal–organic structure MOF-5.

4. Materials and Methods

4.1. Materials

The reactives HCl and AgNO3 were purchased from Aldrich. The gases used for the textural sorption analysis of clinoptilolites (natural and modified) were ultra-high purity CO2, N2 and He gases (>99.999%, INFRA Corp., Puebla City, Mexico).

4.2. Methodology of Dealumination of Clinoptilolite

The zeolites employed in this work were natural zeolites found in Mexico. These samples are from Atzinco, Puebla, Mexico. The ATN tag corresponds to a natural sample that has not undergone any chemical treatment. Modified clinoptilolite samples ATH1, ATH2, ATH3 and ATH4 were prepared from the ATN precursor and treated different times (1, 2, 3 or 4) with 0.01 N solutions of the HCl at 50 °C for 6 h. This chemical treatment provokes impurities elimination and the exchange of protons in the polyvalent cations places. This alteration technique is composed by a sequence of washing cycles of the samples of natural clinoptilolite (mesh 60–80) with HCl diluted. Then, they were subject to prolonged rinses with deionized water, using AgNO3 as indicator. A specimen named ATH1 was prepared from ATN, after the natural zeolite was treated once with HCl (H1 labeling follows the number of washings with HCl). Similarly, a sample branded as ATH2 resulted from ATN after treating twice the natural zeolite with HCl.

4.3. Experimental Measurement Techniques

4.3.1. XRD

The crystalline phases were identified by using the powder X-ray diffraction (XRD) technique. The equipment utilized was diffractometer Bruker model D8 Discover (Bruker, Co., Billerica, MA, USA). This technique employed nickel-filtered Cu Kα (λ = 0.154 nm) radiation operated at 40 kV and 30 mA. The samples were pulverized in an agate mortar and placed in a sample holder, compressing the powder until it was perfectly compact. The crystalline phases were classified by means of the 2013 PDF4+ archive of the International Center for Diffraction Data (ICDD). The quantification of the percentages of the phases presents in ATN, was performed by the High Score Plus 3.0e computer program (Version 5.1, Malvern Panalytical, Cambridge, UK), properly handling the diffraction patterns (XRD) and atomic coordinates of the present phases [26].

4.3.2. SEM

Zeolites images were acquired with a JEOL, model JSM-7800F (JEOL USA, Inc., Peabody, MA, USA) high-resolution scanning electron microscope at 5 kV. The zeolites were placed on aluminum stub holders and successively coated with Au using a sputtering coater.

4.3.3. EDS

An elemental chemical analysis by X-ray energy dispersion (EDS) was performed on each of the zeolites, using an Oxford INCA energy 250+ model probe, with a resolution of 137 eV and a 10 mm2 detector.

4.3.4. N2 Adsorption

The studies of N2 adsorption were obtained at 76.7 K. This temperature is the boiling point of liquid N2 at the 2200-m altitude of Mexico City. The isotherms were measured in an automatic volumetric adsorption equipment (Quantachrome AutoSorb1LC).
N2 adsorption isotherms were determined at relative pressures from p/p0 = 10−6 to 0.995. The pressure of saturation, p0, was uninterruptedly recorded during the adsorption–desorption process. Powder particle sizes, previously mesh 60–80, were obtained from all zeolites studied. Before the sorption experimentations, each portion of zeolites was exposed at 623 K during 15 h at a pressure lesser than 10−6 Torr. The relevant equations used for the calculation of the surfaces were: The Brunauer–Emmett-Teller (BET), Langmuir, and t-plots (external) in the range of p/p0 from 0.01 to 0.25. The total pore volume, V was calculated by using the Gurvitsch rule based on the volume adsorbed at the relative pressure p/p0 = 0.95, calculated as liquid volume.

4.4. Adsorption of CO2

Chromatographic CO2 adsorption tests were made in a GC-14A Shimadzu gas chromatograph (Shimadzu Co KK, Kyoto, Japan) prepared with a thermal conductivity detector. The glass chromatographic columns (i.d. = 5 mm, length = 50 cm) were filled with zeolites (60–80 mesh).
The incumbent GC column was maintained at a certain temperature while adsorbate pulses of different intensities were injected. The elution chromatogram of each pulse was recorded till the recorder pen got once more the baseline. Previously to the adsorption experimentation, the adsorbents were exposed in situ to a flow of He carrier gas at 573 K for 8 h.
The gas in pure form was injected to measure their corresponding retention time inside the appropriate adsorbent column. The dynamic adsorption method of gas chromatography is the technique pursued in this research to acquire the chromatographic peaks. The isosteric enthalpy of adsorption of CO2 were calculated through Clausius–Clapeyron equation, by using sorption data monitors at two temperatures.

4.5. Calculation Methods

The adsorption mechanisms of the zeolites studied in this work have been previously reported [15,16,17,25]. The microporosity of the natural zeolites used in this work was studied with the Dubinin–Polany theory, in the range of p/p0 = [10−6–0.2] and with high resolution t plots in the range of p/p0 = [10−6–0.8]. The Dubinin–Astakhov (D-A) method was used, which is based on the following equation:
W = W 0 exp [ ( A β E 0 ) n ]
where W is the microporous volume that has been filled at a given p/p0 and W0 is the total volume of the microporous space. In turn -A = RT ln(p0/p) is the negative of the isothermal molar Gibbs free energy. β is called a scaling factor and E0 is the characteristic energy of the adsorption system. This term reflects the influence of the nature of the substrate on the adsorbed amount.
The experimental data of the adsorption of CO2 in the zeolites were treated by the adsorption models of Freundlich and Langmuir in their linear form. The Freundlich adsorption equation can be written as:
a = KF + p1/n
where a is the adsorbed amount (mmol g−1), KF is the Freundlich adsorption constant, and n is an exponential factor. From gas adsorption data at low pressures, it is possible to evaluate the Henry constants (KH) at different temperatures for the series of adsorbent–adsorptive pairs employed in this work according to the following expression [27]:
KH = limp→0 (a/amp)
where a represents the amount adsorbed on the solid walls at pressure p, while am is the monolayer capacity evaluated from the Langmuir equation:
θ = a/am = Kp/1 + Kp
where K am = KH, is something that can be tested graphically by plotting 1/a versus 1/p:
1/a = 1/am + 1/amKp
The enthalpy isosteric of adsorption, -qst (kJ mol−1), at different adsorbate loadings can be evaluated from the adsorption isotherms data through a Clausius–Clapeyron type equation [27]:
[∂ ln p/∂T]a = -qst a/RT2
where p and T are the equilibrium pressure and temperature at a given adsorbate loading (a).

5. Conclusions

In this work, the experimental CO2 data were obtained by IGC. Afterwards, the data were treated through the equations of Freundlich and Langmuir. Meanwhile, the degree of interaction between CO2 and the clinoptilolite samples studied was analyzed with the isosteric enthalpy of adsorption. This enthalpy was studied by applying the Clausius–Clapeyron equation.
From the initial results which establishes the following sequence: ATH1 > ATH2 > ATH4 > ATN > ATH3. The initial sharp decrease in -qst may be attribute to the site heterogeneity. Though, a constant gentle reduction in -qst is anticipated at higher adsorbate loadings due to repellent interactions between adsorbed CO2 molecules. Previously, the nanoporosity of natural clinoptilolite zeolites from different locations in Atzinco, Puebla, Mexico was studied through HRADS adsorption of N2. Laterally, these natural zeolites have been conveniently characterized by XRD, EDS, and SEM. The quantitative determination showed that natural tuff contains Ca-Clinoptilolite (88.76%) >> Montmorillonite (11.11%) >> quartz (0.13%). The SiO2/Al2O3 molar ratio for the starting material is 7.75; after acid treatment, this ratio forms a jumping curve: ATH4 > ATH2 > ATN > ATH3 > ATH1. SEM images of these zeolites indicate the presence of crystals with different sizes and morphology, protruding anhedral crystals and lamellar pores. Porosity studies indicate the formation of type VI adsorption isotherms for ATN zeolite, while for the ATH zeolites it describes by mixed type I–IV isotherms with hysteresis cycles of type H3, which are characteristic of lamellar pores. The Langmuir specific surface area (ASL) is: ATN > ATH2 > ATH4 > ATH3 > ATH1. While the VΣ values are as follows: ATN > ATH4 > ATH3 > ATH1 > ATH2.

Author Contributions

Conceptualization, M.A.H.; and R.P.; methodology, V.P., K.M.A.; validation, M.A.H., V.P. and E.R.; formal analysis, R.P.; investigation, M.d.l.A.V.; resources, M.T.; writing—original draft preparation, M.A.H.; writing—review and editing, V.P.; visualization, G.I.H.; supervision, V.P.; project administration, J.D.S. and L.A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the VIEP and the Academic Body “Investigación en zeolitas”, CA-95 (PROMEP-SEP).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hernandez, M.A.; Pestryakov, A.; Portillo, R.; Salgado, M.A.; Rojas, F.; Rubio, E.; Ruiz, S.; Petranovskii, V. CO2 Sequestration by Natural Zeolite for Greenhouse Effect Control. Procedia Chem. 2015, 15, 33–41. [Google Scholar] [CrossRef] [Green Version]
  2. Maurin, G.; Belmabkhout, Y.; Pirngruber, G.; Gaberova, L.; Llewellyn, P. CO2 adsorption in LiY and NaY at high temperature: Molecular simulations compared to experiments. Adsorption 2007, 13, 453–460. [Google Scholar] [CrossRef]
  3. Belmabkhout, Y.; Sayari, A. Effect of pore expansion and amine functionalization of mesoporous silica on CO2 adsorption over a wide range of conditions. Adsorption 2009, 15, 318–328. [Google Scholar] [CrossRef]
  4. Database of Zeolite Structures. Available online: http://www.iza-structure.org/databases/ (accessed on 4 November 2022).
  5. Tsitsihvili, G.V.; Andronikashvili, T.G.; Kirov, G.N.; Filizova, L.D. Natural Zeolites; Ellis Horwood Ltd.: Brighton, UK, 1992; pp. 100–120. [Google Scholar]
  6. Rouquerol, F.; Rouquerol, J. Adsorption by Powders and Porous Solids: Principles, Methodology and Applications, 2nd ed.; Academic Press: Amsterdam, The Netherlands, 2014; pp. 626–632. [Google Scholar] [CrossRef]
  7. Hutson, N.D.; Yang, R.T. Theoretical Basis for the Dubinin-Radushkevitch (D-R) Adsorption Isotherm Equation. Adsorption 1997, 3, 189–195. [Google Scholar] [CrossRef]
  8. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC technical report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  9. Diaz, E.; Ordoñez, S.; Vega, A.; Coca, J. Adsorption characterisation of different volatile organic over alumina, zeolites and activated carbon using inverse gas chromatography. J. Chromatogr. A 2004, 1049, 139–146. [Google Scholar] [CrossRef]
  10. Gürkan, R.; Ceylan, H.; Şahan, T.; Savran, A.; Kubilay, Ş. Evaluation of the Thermodynamic Parameters for the Adsorption of Some Hydrocarbons on Chemically Treated-Bentonites by Inverse Gas Chromatography. Adsorption 2005, 11, 603–612. [Google Scholar] [CrossRef]
  11. Balard, H.; Papirer, E. Characterization and modificatio of fillers for paints and coatings. Prog. Org. Coat. 1993, 22, 1–17. [Google Scholar] [CrossRef]
  12. Bilgic, C.; Aşkınc, A. Evaluation of the thermodynamic parameters for the adsorption of some hydrocarbons on alumina and molecular sieves 3A and 5A by inverse gas chromatography. J. Chromatogr. A 2003, 1006, 281–286. [Google Scholar] [CrossRef]
  13. Jagiello, J.; Bandosz, T.J.; Schwarz, J.A. Application of inversa gas chromatography at infinite dilution to study the effects of oxidation of activated carbons. Carbon 2002, 30, 63–69. [Google Scholar] [CrossRef]
  14. Chen, S.G.; Yang, R.T. Theoretical basis for the potential theory adsorption isotherms. The Dubinin-Radushkevich and Dubinin-Astakhov equations. Langmuir 1994, 10, 4244–4249. [Google Scholar] [CrossRef]
  15. Quiroz, K.; Hernandez, M.A.; Portillo, R.; Rojas, F.; Rubio, E.; Petranovskii, V. Comparación de la capacidad de secuestro de CO2 en Clinoptilolita, Epistilbita, Erionita, Mordenita y Caolinita como constituyentes del suelo. Superf. Vacío 2016, 29, 55–61. [Google Scholar]
  16. Hernandez, M.A.; Portillo, R.; Salgado, M.A.; Rojas, F.; Petranovskii, V.; Pérez, G.; Salas, R. Comparación de la capacidad de adsorción de CO2 en clinoptilolitas naturales y tratadas químicamente. Superf. Vacío 2010, 23, 67–72. [Google Scholar]
  17. Garcia, R.; Hernández, M.A.; Portillo, R.; Petranovskii, V.; Rubio, E. Adsorción irreversible de CO2, H2 y CH4 en zeolitas naturales de poro angosto. Rev. Int. Contam. Ambient. 2017, 34, 685–696. [Google Scholar] [CrossRef]
  18. 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. [Google Scholar] [CrossRef] [PubMed]
  19. Wong-Ng, W.; Kaduk, J.; Huang, Q.; Espinal, L.; Li, L.; Burress, J. Investigation of NaY Zeolite with adsorbed CO2 by neutron powder diffraction. Microporous Mesoporous Mater. 2013, 172, 95–104. [Google Scholar] [CrossRef]
  20. Park, H.J.; Suh, M.P. Enhanced isosteric heat, selectivity, and uptake capacity of CO2adsorption in a metal-organic framework by impregnated metal ions. Chem. Sci. 2012, 4, 685–690. [Google Scholar] [CrossRef]
  21. Treacy, M.M.J.; Higgins, J.B. Collection of Simulated XRD Powder Patterns for Zeolites, 4th ed.; Elsevier: Amsterdam, The Netherlands, 2001; pp. 49–50. [Google Scholar] [CrossRef]
  22. Dávalos, O.G.; Nieto, Á.F.; Alaniz, S.A.; Martinez, E.; Ramirez, E. Cenozoic stratigraphy of the Tehuacan region and its relationship with the northern sector of the Oaxaca fault, Cenozoic stratigraphy of the Tehuacan region and its relationship with the northern sector of the Oaxaca fault. Rev. Mex. Cienc. Geol. 2007, 24, 197–215. [Google Scholar]
  23. Sprynskyy, M.; Golembiewski, R.; Trykowski, G.; Buszewski, B. Heterogeneity and hierarchy of clinoptilolite porosity. J. Phys. Chem. Solids 2010, 71, 1269–1277. [Google Scholar] [CrossRef]
  24. Pour, A.A.; Sharifnia, S.; Salehi, R.N.; Ghodrati, M. Performance evaluation of clinoptilolite and 13X zeolites in CO2 separation from CO2/CH4 mixture. J. Nat. Gas Sci. Eng. 2015, 26, 1246–1253. [Google Scholar] [CrossRef]
  25. Hernández, M.A.; Corona, L.; González, A.I.; Rojas, F.; Lara, V.H.; Silva, F. Quantitative Study of the Adsorption of Aromatic Hydrocarbons (Benzene, Toluene, and p-Xylene) on Dealuminated Clinoptilolites. Ind. Eng. Chem. Res. 2005, 44, 2908–2916. [Google Scholar] [CrossRef]
  26. Degen, T.; Sadki, M.; Bron, E.; König, U.; Nénert, G. The High Score suite. Powder Diffr. 2014, 29, S13–S18. [Google Scholar] [CrossRef] [Green Version]
  27. Thang, H.V.; Grajciar, L.; Nachtigall, P.; Bludský, O.; Areán, C.O.; Frýdová, E.; Bulánek, R. Adsorption of CO2 in FAU zeolites: Effect of zeolite composition. Catal. Today 2014, 227, 50–56. [Google Scholar] [CrossRef]
Figure 1. 3D simulation clinoptilolite structure, that schematically displays Si and O atoms (yellow and red balls, respectively), and the inner surface of the channels (light-blue surface).
Figure 1. 3D simulation clinoptilolite structure, that schematically displays Si and O atoms (yellow and red balls, respectively), and the inner surface of the channels (light-blue surface).
Separations 10 00238 g001
Figure 2. XRD patterns of clinoptilolite zeolite from San Juan Atzinco, Puebla, Mexico (ATN), % of each of the crystalline phases.
Figure 2. XRD patterns of clinoptilolite zeolite from San Juan Atzinco, Puebla, Mexico (ATN), % of each of the crystalline phases.
Separations 10 00238 g002
Figure 3. XRD results of clinoptilolite zeolites at: (a) 2θ = 5–20°, (b) 2θ = 20–35°, (c) 35–50° and (d) 2θ = 5–50°.
Figure 3. XRD results of clinoptilolite zeolites at: (a) 2θ = 5–20°, (b) 2θ = 20–35°, (c) 35–50° and (d) 2θ = 5–50°.
Separations 10 00238 g003
Figure 4. SEM micrographs of natural Ca-Clinoptilolite ATN zeolite from San Juan Atzinco, Puebla, Mexico with different approaches: (a) 15,000× and (b) 40,000×.
Figure 4. SEM micrographs of natural Ca-Clinoptilolite ATN zeolite from San Juan Atzinco, Puebla, Mexico with different approaches: (a) 15,000× and (b) 40,000×.
Separations 10 00238 g004
Figure 5. N2 adsorption isotherms at 77 K on Clinoptilolite-zeolite in (a) logarithmic p/p0 scale and (b) normal scale.
Figure 5. N2 adsorption isotherms at 77 K on Clinoptilolite-zeolite in (a) logarithmic p/p0 scale and (b) normal scale.
Separations 10 00238 g005
Figure 6. Nanopore size distribution of Ca-Clinoptilolite zeolites from Dubinin–Astakhov equation.
Figure 6. Nanopore size distribution of Ca-Clinoptilolite zeolites from Dubinin–Astakhov equation.
Separations 10 00238 g006
Figure 7. PSDs of clinoptilolite-zeolites calculated from BJH method.
Figure 7. PSDs of clinoptilolite-zeolites calculated from BJH method.
Separations 10 00238 g007
Figure 8. Adsorption isotherms of CO2 on ATN and ATH clinoptilolite zeolites at distinct temperatures, (a) 433 K, (b) 473 K, (c) 513 K, (d) 543 K and (e) 573 K.
Figure 8. Adsorption isotherms of CO2 on ATN and ATH clinoptilolite zeolites at distinct temperatures, (a) 433 K, (b) 473 K, (c) 513 K, (d) 543 K and (e) 573 K.
Separations 10 00238 g008
Figure 9. Adsorption isosteric enthalpy of CO2 on ATN and ATH clinoptilolites.
Figure 9. Adsorption isosteric enthalpy of CO2 on ATN and ATH clinoptilolites.
Separations 10 00238 g009
Table 1. Crystalline phases, XRD.
Table 1. Crystalline phases, XRD.
SampleCard Number%Name
(Na, Ca)0.3(Al, Mg)2Si4O10(OH)2 x H2O00-003-001511.110Montmorillonite (Bentonite)
Ca3.16 Si36O72 (OH2)21.8001-070-185988.761Ca-Clinoptilolite
SiO201-085-07950.128Quarzt
Table 2. Samples chemical composition (mass %), based on EDS data.
Table 2. Samples chemical composition (mass %), based on EDS data.
SampleSiO2Al2O3Fe2O3CaOMgOK2OTiO2FeOSi/Al
ATZN59.307.651.546.325.992.540.601.397.75
ATH163.5969.1072.7162.2384.3645.830.8671.2226.983
ATH262.4847.9353.462.5186.0284.9390.7841.5567.874
ATH363.4259.8623.0882.9524.3976.1670.6171.3896.431
ATH461.8217.8223.9462.5045.7324.9140.8671.7757.903
Table 3. Textural parameters of clinoptilolites from Atzinco, Puebla.
Table 3. Textural parameters of clinoptilolites from Atzinco, Puebla.
SampleASL
m2/g
ASB
m2/g
ASt
m2/g
CBZone p/p0
BET
V
cm3/g
W0t
cm3/g
ATN40.5026.1715.54950.054–0.2890.0750.010
ATH112.399.195.912−3670.010–0.2100.01570.0058
ATH231.0623.3917.032870.010–0.2140.01560.0058
ATH315.8810.346.6826100.01–0.3140.02280.0016
ATH430.2920.5118.32−3190.01–0.3140.03560.0012
ASL, is the specific surface area Langmuir; ASL is the specific surface area BET; ASt is the external surface area, t-plot; CB is the BET constant; p/p0 is the range used for BET plot, VΣ is the volume adsorbed close to saturation (p/p0 = 0.95), calculated as volume of liquid (Gursvitch rule) and W0t is micropore volume, t-plot.
Table 4. Clinoptilolite zeolites micropores parameters according to the Dubinin–Astakhov equation.
Table 4. Clinoptilolite zeolites micropores parameters according to the Dubinin–Astakhov equation.
SampleE
Kj mol−1
W0
cm3 g−1
nDp
nm
Al2O3M+
ATN2.560.0210.9507.656.32 MgO
ATH13.160.0110.8909.1075.83 K2O
ATH22.850.0210.9207.9356.028 MgO
ATH32.730.0110.9309.8626.167 K2O
ATH45.780.012.700.7907.8225.732 MgO
Table 5. Henry (KH, mmol g−1 x mmHg−1), Freundlich (KF, mmol g−1 × mmHg−1) and Langmuir (am, mmol g−1) parameters for the adsorption of CO2 on modified clinoptilolite zeolites.
Table 5. Henry (KH, mmol g−1 x mmHg−1), Freundlich (KF, mmol g−1 × mmHg−1) and Langmuir (am, mmol g−1) parameters for the adsorption of CO2 on modified clinoptilolite zeolites.
T/KSampleKF 103nRFKH 103amRL
573ATN6.41.2900.9954.90.2070.999
ATH111.91.4500.9918.80.2100.999
ATH211.71.3800.9959.20.2230.999
ATH39.41.3430.9977.30.2180.998
ATH411.91.3950.9969.50.2050.998
543ATN8.91.4380.9906.30.1740.999
ATH18.41.3490.9965.90.2470.999
ATH212.81.4110.994100.2230.999
ATH38.41.3490.9967.90.2050.999
ATH410.71.4250.99310.20.2090.999
513ATN9.31.4500.9896.60.1730.999
ATH112.61.4080.9949.80.2230.999
ATH214.91.4810.99311.40.2120.998
ATH3131.5140.9899.20.1970.999
ATH411.71.3830.99390.2300.999
473ATN10.21.3830.9937.60.1620.999
ATH115.81.5070.99111.70.2200.999
ATH219.21.5930.99615.80.1850.995
ATH314.51.5260.98911.80.1680.999
ATH415.81.4620.98713.50.1930.999
433ATN7.41.2120.9916.10.3120.999
ATH117.21.5070.99414.70.1890.996
ATH217.91.5210.98514.60.2000.999
ATH320.11.6380.989170.1680.999
ATH417.71.4690.99315.50.1980.998
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

Hernandez, M.A.; Hernandez, G.I.; Portillo, R.; Rubio, E.; Petranovskii, V.; Alvarez, K.M.; Velasco, M.d.l.A.; Santamaría, J.D.; Tornero, M.; Paniagua, L.A. CO2 Adsorption on Natural Zeolites from Puebla, México, by Inverse Gas Chromatography. Separations 2023, 10, 238. https://doi.org/10.3390/separations10040238

AMA Style

Hernandez MA, Hernandez GI, Portillo R, Rubio E, Petranovskii V, Alvarez KM, Velasco MdlA, Santamaría JD, Tornero M, Paniagua LA. CO2 Adsorption on Natural Zeolites from Puebla, México, by Inverse Gas Chromatography. Separations. 2023; 10(4):238. https://doi.org/10.3390/separations10040238

Chicago/Turabian Style

Hernandez, Miguel Angel, Gabriela Itzel Hernandez, Roberto Portillo, Efraín Rubio, Vitalii Petranovskii, Karin Montserrat Alvarez, Ma de los Angeles Velasco, Juana Deisy Santamaría, Mario Tornero, and Laura Alicia Paniagua. 2023. "CO2 Adsorption on Natural Zeolites from Puebla, México, by Inverse Gas Chromatography" Separations 10, no. 4: 238. https://doi.org/10.3390/separations10040238

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