Accès libre

Evaluation of CO2 adsorption capacity with a nano-CaO synthesized by chemical combustion/ball milling

À propos de cet article

Citez

Introduction

The accelerated increase in atmospheric CO2 concentration due to use of energy resources, such as the combustion of fossil fuels, is a cause of climate change, a vital global problem [1]. Thus, large-scale mitigation of CO2 emissions represents an urgent need; great efforts have been made to develop novel CO2-capture technologies considering efficiency, reductions in operating and maintenance costs, and retrofitting of power plants, among other characteristics [2]. Amine solvent adsorption, cryogenic methods, adsorption processes over solid adsorbents, and separation with porous membranes [3, 4, 5, 6] are among the proposed methods that offer economic and environmental advantages; however, some of them have several important drawbacks in their use and application on a large scale [7].

CO2 capture by adsorption in solid adsorbents is widely recognized as one of the most promising technologies to be used on a large scale, due to several factors, including its low operating costs and ease of application [8]. With this emphasis, great attention has been directed to the generation of highly promising adsorbents with improved properties to efficiently capture CO2, such as activated carbons, zeolites, metal–organic frameworks, and metal oxides, among others [9]. Metal oxides have received particular attention, and it is important to note that calcium oxide (CaO) has been regarded as the most promising candidate for CO2 capture, particularly structured as CaO nanoadsorbents [10, 11, 12, 13, 14, 15], due to its long durability, good mechanical strength, high reactivity with CO2, and low costs; indeed, it can be obtained from several sources and using various methods [16, 17, 18]. However, more research is needed before the CaO-based adsorbent can be placed in large-scale industrial use.

To improve CaO stability and functionality for CO2 capture, different strategies have been examined [19]. One of them is the reduction of the particle size to the nanometer scale (<100 nm), looking for higher contact and reaction efficiencies; for this goal, mechanical milling is quite useful [20]. Data about the use of CaO adsorbents for CO2 under typical flue gas conditions are limited [21, 22]. However, based on a theoretical calculation, Rashidi et al. [23] reported than 1 g of CaO can capture 786 mg of CO2, whereas activated carbon typically captures 80 mg of CO2/g. In this regard, CaO-based adsorbents are suitable materials to guarantee the industrial competitiveness of CO2-capture technology and environmental applications at a relatively low cost. In a recent study [21], a novel porous nano-CaO adsorbent obtained by chemical combustion and ball milling was reported. This fast synthesis yielded interesting results because the ball milling treatment led to a high surface area, a low average crystalline size, and thus to a nanostructured nature. Structural and textural characteristics of these CaO nanoparticles indicated that they were highly reactive toward the adsorption of CO2 molecules; therefore, interest remained about the influence of the adsorption parameters.

The main purpose of this study was to test the CO2 adsorption behavior of the nano-CaO adsorbent, which was synthesized by the chemical combustion/ball milling method, and subjecting the adsorption process to different conditions. Kinetic, isotherm, and thermodynamic parameters were evaluated at moderate temperatures (25°C, 50°C, and 75°C) and supply pressures (1–15 atm). The properties of the nano-CaO adsorbent in terms of CO2 desorption, stability, and regeneration ability were also analyzed. The study was focused to provide a better understanding of the adsorption mechanism and offer a useful approach to predict the adsorption behavior of this adsorbent.

Experimental
Reagents and materials

The analytical-grade reagents used for the chemical combustion synthesis of nano-CaO adsorbent were calcium nitrate tetrahydrate (Merck ACS, 98.0% purity) as the oxidizing agent and urea (Sigma Aldrich ACS, 99.9% purity) as the chemical fuel; these reagents were used without further purification.

Synthesis of the nano-CaO adsorbent

The nano-CaO adsorbent was prepared via a typical chemical combustion followed by a ball milling process [21]. The redox mixture contained urea as the chemical fuel and calcium nitrate tetrahydrate as the oxidizing agent, according to the following reaction: 2[Ca(NO3)24H2O](s)+4CH4N2O(s)+O2(g)2CaO(S)+16H2O(g)+4CO2(g)+6N2(g) \matrix{{2{{[{\rm{Ca}}{{({\rm{N}}{{\rm{O}}_3})}_2} \cdot 4{{\rm{H}}_2}{\rm{O}}]}_{({\rm{s}})}} + 4{\rm{C}}{{\rm{H}}_4}{{\rm{N}}_2}{{\rm{O}}_{({\rm{s}})}} + {{\rm{O}}_{2({\rm{g}})}}} \hfill \cr {\to 2{\rm{Ca}}{{\rm{O}}_{({\rm{S}})}} + 16{{\rm{H}}_2}{\rm{O}}_{({\rm{g}})} + 4{\rm{C}}{{\rm{O}}_{2({\rm{g}})}} + 6{{\rm{N}}_{2({\rm{g}})}}} \hfill \cr}

A 2:1 molar ratio of urea: calcium nitrate and a minimum volume of distilled water were thoroughly mixed in a porcelain crucible, so as to obtain a homogeneous mixture. This slurry was warmed to almost dryness and then 1 mL of distilled water was added. Finally, the resulting product was heated for 5 min at 800°C. The process was rich in fuel, which facilitated the formation of CaO because enough energy was available during combustion. The chemical combustion method has been traditionally used to prepare several metal oxides and other materials with fluffy structure, large surface areas, high porosity, and high yield [24]. The dried powders derived from the chemical combustion were subjected to mechanical ball milling for 2.5 h under argon atmosphere with a Spex-type mechanical ball mill, with stainless steel balls of 1 mm diameter, and a ball-to-powder weight ratio of 6:1. The dry ground CaO was the starting material for the study of CO2 adsorption.

Adsorbent characterization

The CaO was characterized using the techniques described below.

X-ray diffraction (XRD) analysis with a SIEMENS D-5000 powder diffractometer in a wide-angle region of 10°–100° in the 2θ scale, and the diffraction pattern was compared to the database in the International Centre for Diffraction Data, Pennsylvania, USA.

Scanning electron microscopy (SEM) with electron diffraction scattering (EDS) analysis (JEOL JSM-5900LV with an Oxford EDS microprocessor) for knowing the morphology of the CaO powders and their elemental semiquantitative composition.

N2 physisorption measurements were performed at −196°C (Belsorp Max Japan Inc. instrument), and the corresponding surface area was determined by the Brunauer-Emmett-Teller (BET) method of multiple points; the total pore volume (VTp), the mean pore radius (Rp), and the pore size distribution were derived using the Barrett-Joyner-Halenda (BJH) method. To obtain the greatest gas adsorption capacity and to release the water and impurities adsorbed, the CaO samples were degassed at 200°C for 3 h under a nitrogen atmosphere before measurements.

CO2 adsorption kinetic experiments

Kinetic adsorption experiments of CO2 onto the nano-CaO adsorbent were conducted at different temperatures (25°C, 50°C, and 75°C) and at room pressure using a 50 mL capacity Parr 4592 stainless steel pressure reactor coupled to a temperature-controlled system. Samples of 10 mg of the nano-CaO adsorbent were exposed to a high-purity (99.98%) dry CO2 flow at 1 atm pressure for 1 min, 3 min, 5 min, 10 min, 15 min, 20 min, 25 min, 30 min, 45 min, 60 min, 90 min, and 180 min. Prior to the CO2 adsorption tests, the adsorbent samples were degassed for 30 min under vacuum at 325°C to remove some possible impurities. Measurements of the CO2 adsorption capacity were performed by thermogravimetric analysis (TGA) with a TGA analyzer (SDT Q600; TA Instruments) coupled with a mass spectrum analyzer (TA Instruments, LLC). For these analyses, the CaO sample (~6 mg) was put in a ceramic cell and then heated from 20°C to 850°C at the rate of 10°C/min, in an inert atmosphere of helium (100 mL/min). The quantity of CO2 (millimoles) captured per gram of adsorbent (millimoles per gram [mmol/g]) was calculated from the TGA calcination profile for different contact times and temperatures based on the weight loss of CO2; the first derivatives of the TGA curves (differential thermal analysis [DTA]) were also computed.

CO2 adsorption isotherms

The adsorption isotherms of CO2 in the porous nano-CaO adsorbent were determined at 25°C, 50°C, and 75°C and under different CO2 gas pressures (1 atm, 5 atm, 10 atm, and 15 atm). The adsorption capacity of CO2 was measured by TGA as for the kinetic tests. Classic Langmuir and Freundlich isotherm models were used to provide a mathematical description of the CO2 adsorption process on the nano-CaO adsorbent, and their respective constants were determined. These isotherm models generally provide information on the CO2 adsorption mechanisms and affinity properties of the adsorbents.

CO2 adsorption thermodynamics

The effect of temperature on the amount of CO2 adsorbed by the nano-CaO adsorbent was studied at 298 K, 323 K, and 348 K. The nano-CaO adsorbent dosage, the flow, and the CO2 pressure were constant. The most important thermodynamic parameters, such as standard enthalpy (ΔH°), standard entropy (ΔS°), standard free energy (ΔG°), and activation energy (Ea), of CO2 adsorption were determined.

Cyclic CO2/nano-CaO adsorption-desorption

The CO2 adsorption-desorption activity of the nano-CaO adsorbent was determined as follows. First, adsorption of CO2 onto nano-CaO adsorbent was carried out at 75°C and 1 atm, and the adsorption capacity qad was measured in these conditions. Then, these CO2/nano-CaO samples were exposed to a helium flow from 20°C to 850°C for eliminating CO2. Thereafter, CO2 was again adsorbed onto the same samples of the nano-CaO adsorbent. Five cycles of adsorption-desorption were achieved.

Results and discussion
Adsorbent characterization

The chemical combustion/ball milling method allowed the preparation of a large amount of voluminous, spongy, and nanocrystalline CaO adsorbent with a large surface area and high porosity. Figure 1 shows the XRD pattern of the adsorbent, in which the CaO is the most important mineral component of the sample (International Centre for Diffraction Data, file: 37-1497). The crystallite size of the CaO particles using the Scherrer equation was 18.61 nm, showing a nanostructured material. In addition, a crystalline structure of CaCO3 appears as a trace component (International Centre for Diffraction Data, file: 05-0586).

Fig. 1

XRD diffraction pattern of the nano-CaO adsorbent.

Note: XRD, X-ray diffraction

The SEM micrograph of the CaO (Figure 2) adsorbent shows an agglomeration of nanoparticles with a well-developed and irregularly shaped porous structure, rough surface, and a small particle size (<10 μm) due to the fragmentation caused by the mechanical ball milling. In this regard, the agglomerates have substantial surface texture, porosity, and big surface area. A porous structure facilitates CO2 access to the active sites and thereby increases the rates of CO2 physisorption, chemisorption, and diffusion. The quantitative microanalysis of the nano-CaO adsorbent using SEM with X-ray energy dispersion only confirmed the presence of calcium and oxygen elements in this sample; a high CaO content in the adsorbent structure favors the conversion of CaO into CaCO3.

Fig. 2

SEM micrograph of Cao adsorbent at 5000× and EDS analysis.

Notes: EDS, electron diffraction scattering; SEM, scanning electron microscopy

N2 physisorption at 77 K showed that the CaO adsorbent had a large BET surface area of 50.73 m2/g (which is indicative of highly developed porosity), a high total pore volume of 0.548 cm3/g, and a narrow and uniform size distribution of the pore radius, which was determined by the BJH method to be 19.24 nm with an ~80% distribution (Figure 3). These improved structural and textural properties all played an important role in capturing CO2 and showed a high CO2 adsorption performance of the adsorbent when it was exposed to a flow of CO2 at different temperatures.

Fig. 3

Determination of pore-size distribution by the BJH method.

Notes: BJH, Barrett-Joyner-Halenda

CO2 adsorption kinetics

The adsorption kinetics describes the adsorption rate of the adsorbate, which governs the residence time of the adsorption reaction. The rate at which adsorption occurs is therefore of paramount importance in the design of an appropriate adsorption system, and thus, understanding the kinetics of adsorption is vital for capturing CO2 because it provides essential information on the possible speed control steps (e.g., chemical reaction, diffusion control, and mass transfer) [25]. In the present study, kinetic models were used to investigate the CO2 adsorption behavior on nano-CaO adsorbent under a range of temperatures and durations, ranging from 1 min to 180 min. The correlation coefficients (R2) determined the conformity between the experimental data and each kinetic model, which were the pseudo-first-order [26], pseudo-second-order [27], Elovich [28], and intraparticle diffusion [29] models, the most widely used for describing adsorption processes. The results obtained are shown in Table 1.

Kinetic parameters evaluated for the process of CO2 adsorption onto the nano-CaO adsorbent at different temperatures

Models Parameters Temperature
25°C 50°C 75°C

Pseudo-first-order model k1 (/min) 0.24 0.58 0.13
qmax (mmol/g) 4.59 3.81 5.27
R2 0.78 0.58 0.74
Pseudo-second-order model k2 (/min) 3.41 1.4 0.20
qmax (mmol/g) 9.02 10.08 11.02
R2 0.9999 0.9996 0.9998
Elovich α (mmol/g) 176.8 122.1 41.2
β (mmol/g) 1.00 1.16 1.40
R2 0.48 0.50 0.59
Intraparticle diffusion

Step 1 (0 < t < 1 min)

Step 2 (1 < t < 10 min)

qe (mmol/g) 7.56 8.33 8.53
kip1 (mmol/g min1/2) 4.36 4.81 4.92
R2 0.87 0.85 0.90
qe (mmol/g) 3.38 3.51 4.64
kip2 (mmol/g min1/2) 0.87 0.90 1.19
R2 0.995 0.94 0.94

Data of Table 1 show that the correlation coefficient (R2) for the pseudo-second-order model is higher than for the other ones, which indicates that this model was the most appropriate for describing the kinetics of CO2 adsorption onto the nano-CaO adsorbent. The corresponding equation is as follows (Figure 4): 1qt=1k2qe2+(1qe)t, {1 \over {{q_t}}} = {1 \over {{k_2}{q_e}^2}} + \left({{1 \over {{q_e}}}} \right)t, where qe and qt are the amounts of CO2 adsorbed (in mmol/g) at equilibrium and at time t (in minutes), respectively, and k2 is the pseudo-second-order adsorption rate constant (/min). According to this model, it is assumed that the rate-limiting step can be chemisorption, which involves valence forces through the sharing or exchange of electrons between adsorbate (gas molecules) and adsorbent [29]. Moreover, the occupation of the velocity adsorption sites is proportional to the number of unoccupied sites, and the adsorption site occupancy rate is proportional to the square of the number of unoccupied sites.

Fig. 4

Kinetic data of CaO adsorption onto nano-CaO adsorbent, at different temperatures and 1 atm pressure.

Notes: Linear plots of the pseudo-second-order model: t/qt = 0.11t + 0.004 (25°C); t/qt = 0.099t + 0.02 (50°C); t/qt = 0.089t + 0.05 (75°C); R2 = 0.9. qt = the amount of CO2 adsorbed at time t; R2, correlation coefficient

For the process CO2/nano-CaO, the maximum adsorption capacity (qe, mmol/g) increases with temperature (25°C, 50°C, and 75°C) at room pressure, whereas the rate constant diminishes with this parameter. This effect suggests a greater driving force of the gas molecules when temperature increases, leading to faster diffusion of CO2 and an increase in CO2 capture. These kinetic results are important regarding cost-effective adsorption technologies because the pollutant CO2 is removed in a relatively short time. For all the studied conditions, CO2 adsorption was observed in the first minute of contact, equilibrium was achieved in approximately 15 min, and no further changes in CO2 adsorption were observed afterward. The increasing of CO2 adsorption with the contact time prior to reaching equilibrium is due to the high availability of the active sites of the adsorbent, and the process slowed in the later stages of contact time due to a decrease in the number of active sites. The experiments for determining other adsorption parameters were carried out with 20 min contact time.

The values of R2 for the intraparticle diffusion model are good enough as well. The equation of this model is the following: qt=KiPt+I, {q_t} = {K_{iP}}\sqrt t + I, where KiP corresponds to the rate constant of each step of the process and I is the intercept that provides an idea about the boundary layer thickness: i.e., the larger the intercept, the greater is the boundary layer effect (in mmol/g). According to this model, the gas molecules may diffuse into the particles when the solid adsorbent is porous enough. The adsorption of the gaseous phase on and in the adsorbent may occur via three main stages: (1) transport of the gas molecules from the volumetric phase to the adsorbent surface; (2) diffusion of the gas molecules from the surface to the adsorbent pores; and (3) diffusion in the interior of the pores.

Data on the adsorption of CO2 by the CaO nanoparticle adsorbent treated according to Eq. (2) showed that it was a multilinear process (Figure 5). The first line corresponds to the instant adsorption from the volumetric phase to the surface of the adsorbent; the second section is attributed to the gradual intraparticle diffusion; and the third line corresponds to the final equilibrium stage [25]. The first stage was very fast and was completed in the first 1 min; the second stage of intraparticle diffusion was reached at 10 min; thereafter, equilibrium was achieved. The value of KiP for the first stage was higher than that of the second one (Table 1). Therefore, the adsorption of CO2 on the highly porous nano-CaO adsorbent was probably not only due to intraparticle diffusion but this mechanism may be a consequence of the adsorption itself.

Fig. 5

Kinetic data of CaO adsorption onto nano-CaO adsorbent, at different temperatures under 1 atm pressure.

Notes: Linear plots correspond to the intraparticle diffusion kinetic model. (1) qt = 7.1√t, qt = 0.73√t+6.2 (R2 = 0.94), and qe = 9.2 (25°C). (2) qt = 7.7√t, qt = 0.86√t+6.7 (R2 = 0.96), and qe = 10.3 (50°C). (3) qt = 7.72√t, qt = 0.86√t+6.7 (R2 = 0.97), and qe = 10.9 (75°C). qe, the amount of CO2 adsorbed at equilibrium; qt = the amount of CO2 adsorbed at time t; R2, correlation coefficient

CO2 adsorption isotherms

Data of qe obtained under different supply pressures (1 atm, 5 atm, 10 atm, and 15 atm) and at different temperatures (25°C, 50°C, and 75°C) were treated using the linear equations of both Langmuir [30] and Freundlich [31] adsorption isotherm models. The results are presented in Table 2, where the higher value of the correlation coefficient (R2) corresponds to the Langmuir model. The linear form of this model is expressed as follows: PCO2qe=(1qmax)PCO2+(1KLqmax) {{{P_{{CO}_2}}} \over {{q_e}}} = \left({{1 \over {{q_{max}}}}} \right){P_{{CO}_2}} + \left({{1 \over {{K_{L\,{q_{max}}}}}}} \right) where PCO2 is the pressure (in atmospheres) of the adsorbed gas CO2, qe is the adsorption capacity of CO2 at equilibrium (mmol/g), qe is the maximum adsorption capacity of CO2 at equilibrium (mmol/g), and KL (/atm) is the Langmuir isotherm constant, which represents the affinity between the adsorbent and adsorbate and may be used to predict whether an adsorption system is favorable or unfavorable.

Summarized linear isotherm parameters for CO2 adsorption on nano-CaO adsorbent at the temperatures of 25°C, 50°C, and 75°C and at pressures of 1 atm, 5 atm, 10 atm, and 15 atm for each temperature

Temperature (°C) Langmuir constants Freundlich constants


qmax (mmol/g) KL (/atm) R2 KF (mmol/g [atm]1/n) 1/n R2

25 10.74 0.48 0.996 0.111 0.049 0.78
50 11.31 0.35 0.998 0.099 0.022 0.75
75 12.55 0.31 0.998 0.091 0.009 0.88

The Langmuir isotherm parameters at 25°C, 50°C, and 75°C were evaluated from the slopes and the intersections of the plots PCO2/qe versus PCO2 (Figure 6). As expected, the CO2 adsorption capacity increased with the CO2 pressure, which is the driving force of the adsorption process. Based on the data of Table 2, both KL and qmax increase with temperature. This indicates that the adsorbent surface was well coated with CO2 molecules as a result of their higher surface affinity by implementing a chemisorption behavior. CO2 must be chemisorbed on fixed surface sites.

Fig. 6

Isotherm data of CaO adsorption onto nano-CaO adsorbent, at different temperatures.

Notes: Linear equations of the Langmuir model: (1) PCO2/qe = 0.093 PCO2 +0.045 (R2 = 0.996) (25°C). (2) PCO2/qe = 0.088 PCO2 +0.031 (R2 = 0.998) (50°C). (3) PCO2/qe = 0.080 PCO2 +0.024 (R2 = 0.998) (75°C). PCO2, pressure of the adsorbed gas CO2; qe is the maximum adsorption capacity of CO2 at equilibrium; R2, correlation coefficient

Moreover, according to the Langmuir model, the surface of the porous nano-CaO adsorbent is homogeneous, and CO2 adsorption occurs in a monolayer of CaO; in this case, the formation of multilayers, as proposed by the Freundlich model, is restricted [32].

The essential characteristics of the Langmuir isotherm can be expressed in terms of a dimensionless constant called the separation factor (RL), which examines the favorable acceptance of the adsorption process [33]. An adsorption process is irreversible if RL =0, favorable if 0 < RL < 1, linear if RL = 1, and unfavorable if RL > 1. The equation for determining the dimensionless RL values is the following: RL=11+KLPCO2 {R_L} = {1 \over {1 + {K_L}{P_{{CO}_2}}}} where KL is the Langmuir constant or the affinity constant (/atm), and PCO2 is the equilibrium pressure of the adsorbed gas (in atmospheres). From the experimental results, the RL terms showed values from 0.67 to 0.12 (25°C), from 0.74 to 0.16 (50°C), and from 0.77 to 0.18 (75°C), for the range of pressure from 1 atm to 15 atm. These values of RL, which are <1, indicate that the conditions of the adsorption process are favorable.

CO2 adsorption thermodynamics

Important factors, such as energy changes in a reaction, should be considered when developing an adsorption system, and thermodynamic concepts are required to achieve this outcome. The thermodynamic parameters provide in-depth information on the inherent energy changes associated with adsorption mechanisms. To evaluate these parameters, the linearized form of the van’t Hoff modified equation is the following: lnKL=ΔH0R(1T)+ΔS0R, \ln {K_L} = - {{\Delta {H^0}} \over R}\left({{1 \over T}} \right) + {{\Delta {S^0}} \over R}, where KL is the Langmuir equilibrium constant; ΔH° and ΔS° are the enthalpy and entropy changes, T is the absolute temperature (in Kelvin), and R is the universal gas constant (8.314 × 10−3 kJ/mol K). The slope and intersection of the straight-line diagram of lnKL versus 1/T allows the determination of the values of ΔH° and ΔS° (Figure 7).

Fig. 7

Thermodynamic data of CaO adsorption onto nano-CaO adsorbent, at different temperatures under 1 atm pressure.

Notes: Linear plot of the van’t Hoff modified equation: ln KL = −3917 (1/T) + 12.2 (R2 = 0.9). KL, the Langmuir isotherm constant; R2, correlation coefficient

The thermodynamic analysis revealed that the adsorption process is endothermic due to the positive value of ΔH° = 32.6 KJ/mol. The magnitude of ΔH° can also be used to distinguish between the mechanisms of physisorption (<20 kJ/mol) and chemisorption (~80 kJ/mol to 200 kJ/mol) [34]. As physisorption arises from relatively weak interactions, such as the van der Waals force, chemisorption involves stronger chemical interactions. According to Moroto-Valer et al. [35], the process of physisorption involves a high surface adsorption of energy and the diffusion of molecules at elevated temperatures. The ΔH° value obtained for CO2 adsorption onto the CaO sample was 32.6 kJ/mol, which is indicative of both physisorption and chemisorption processes. The ΔS° value was 0.102 kJ/mol K, and the positive value reflects the affinity of CO2 for the nano-CaO adsorbent and suggests a high disorder of the adsorbate molecules during the fixation of the CO2 molecules onto the active sites of CaO. According to Pighini et al. [36], entropy change is related to the mobility of CO2 and to the coverage on the adsorbent; a high entropy value must be related to chemisorption.

The Gibbs free energy change (ΔG°) is an important criterion for spontaneity, and both energetic and entropy factors must be considered for its evaluation; according to the following equation: ΔG=ΔHTΔS. \Delta {G^ \circ} = \Delta {H^ \circ} - T\Delta {S^ \circ}.

The results of ΔG° were of a positive sign, which implies a nonspontaneous adsorption reaction [37]; moreover, these values diminish with temperature (3.2 KJ/mol, 1.7 KJ/mol, and 0.7 KJ/mol at 298 K, 232 K, and 348 K, respectively), which suggests that the adsorption of CO2 on the CaO-based nanoadsorbent is more favored at higher temperatures.

The adsorption activation energy (Ea) was calculated by using the Arrhenius linear equation: lnKv=EaR(1T)+lnA \ln {K_v} = - {{{E_a}} \over R}\left({{1 \over T}} \right) + lnA where kv is the velocity constant, A is the independent temperature factor (millimoles of substrate per milligram per minute [mmol/mg min]), R is the universal gas constant (8.314 × 10−3 kJ/mol K), T is the temperature (in Kelvin), and Ea is the activation energy of adsorption (in kilojoules per mole [kJ/mol]). This is the minimum energy that must be overcome by the adsorbate molecules and it defines the velocity of the process [38].

The values of the velocity constants of the pseudo-second-order kinetic model at 298 K, 323 K, and 348 K (Table 1) were incorporated into Eq. (7) as a function of 1/T (Figure 8). The values of the slope and the ln kV = 0 intersection of that plot allowed us to obtain Ea = −73.4 KJ/mol and ln A = ~26.9. Otherwise, considering the experimental data of kV at 298 K and 348 K and the Arrhenius equation, in which the factor A is eliminated, we get the following expression: ln(Kv1Kv2)=EaR(1T11T2) ln\left({{{{K_{v1}}} \over {{K_{v2}}}}} \right) = - {{{E_a}} \over R}\left({{1 \over {{T_1}}} - {1 \over {{T_2}}}} \right)

Fig. 8

Thermodynamic data of CaO adsorption onto nano-CaO adsorbent, at different temperatures under 1 atm pressure.

Notes: Linear plot of the Arrhenius equation: ln kV = 8830 (1/T) – 26.9 (R2 = 0.93). kv, the velocity constant; R2, correlation coefficient

The calculated value of Ea = −74.3 KJ/mol was virtually the same as that obtained from Eq. (7). The value of Ea (~80 KJ/mol) confirms that the adsorption mechanism is chemisorption. Typically, physisorption requires low Ea values, whereas chemisorption is characterized by higher values [39]. In general and based on the signs and magnitudes of the thermodynamic parameters obtained in the present study, it can be established that the adsorption of CO2 on the porous nano-CaO adsorbent is mainly produced by chemical adsorption through a thermodynamically nonspontaneous process, and the adsorption of CO2 molecules is positively temperature dependent. These data provide valuable information for the design of improved adsorption schemes to capture CO2 using a new nanoadsorbent based on CaO.

Mechanism of CO2 adsorption onto the nano-CaO adsorbent

CO2 capture technologies require the understanding of the adsorption mechanism to enable use of an adsorbent at the industrial level. According to the literature, a gas-solid carbonation reaction occurs when CaO is directly combined with CO2 [40]. In the CaO-CO2 reaction, the CO2 molecules can interact with the CaO surface, and when this occurs, carbonation directly follows. The chemical reaction may progress until the complete formation of the carbonate thin layer on the surface of the CaO particles, but this layer may block the access of CO2 to the internal active CaO sites of the particles. For chemisorption, the CO2 molecule must diffuse into the pores of the adsorbent and interact with the internal surface of the adsorbent [41]; it has also been shown that reducing the crystallinity of the adsorbent favors diffusion and promotes recharging [42].

In this study, the crystalline structure of the adsorbent, based on CaO, and both the particle and crystal sizes were reduced by the ball milling treatment. This caused a beneficial effect of more conversion from CaO to CaCO3, attributed to the CaO nanoadsorbent’s fluffy structure, significant surface area, and large pore volume, which allows CO2 molecules to easily diffuse into the pore structure and reach deeper active sites, which leads to maximum CO2 capture.

In this framework, the nanostructured nature of CaO caused reactivity and increased adsorption effects involving many active surface sites available for CO2 adsorption. This also allowed the introduction of a large amount of CO2 into its mesopores in a well-dispersed manner that favors the gas-solid diffusion reaction and results in increased CO2 capture. As observed in our results, a mesoporous nano-CaO adsorbent, with a mean pore radius of 19.24 nm at approximately 80%, was obtained. In this way, we improved intraparticle diffusion in the CO2 capture process in the porous CaO because the kinetic radius was sufficient for the diffusion of CO2 molecules within the CaO porous structure when considering a CO2 molecular size of 0.33 nm. The carbonation reaction in the micro-pores is hindered by a limited space that restricts the growth of CaCO3. For CO2 capture with CaO-based adsorbents, mesopores (2–50 nm in diameter) are expected to play an important role based on previous studies on carbonation. Furthermore, these pore types are the dominant contributors to the carbonation reaction and have a high surface area and high pore volume that accommodates the bulky CaCO3 product [43]. The chemical adsorption for the whole process in this material has also been confirmed. Yet, to the best of our knowledge, the physical adsorption process is also controlled by the pore characteristics of the adsorbent, and the micropores (dp <2 nm) create more selective CO2 adsorption. According to Lee and Park [44], the gas adsorption potential reaches a maximum at the subnanometer scale, when the gas molecules begin to overlap with reducing pore size. This results in a much higher bonding energy emanating from the formation of deep potential fields from the neighboring walls, which leads to an increase in the surface area. In our case, narrow pores of the prepared nano-CaO adsorbent were approximately 15% of the total, which were also responsible for the physical capture of CO2.

Finally, it is possible that the most appropriate reaction mechanism in this solid-gas system mainly occurred by chemisorption, with a carbonation reaction on the surface amid a favorable pore morphology of the CaO adsorbent obtained after treatment by ball milling. This was clearly verified by XRD, Fourier transform infrared spectroscopy (FTIR), and X-ray photoelectron spectroscopy (XPS) analysis. However, a very small fraction of physisorption, as corroborated by FTIR [21], was present along with the chemisorption process. In general, all the previously mentioned analysis results confirm the remarkable contribution of chemical adsorption throughout the CO2 adsorption process in the nano-CaO adsorbent. One of the innovative aspects of this work was the favorable capacity of CO2 capture in this adsorbent at moderate temperature and pressure.

In postcombustion processes from large stationary sources, the flow gas is released at temperatures in the range of 40–160°C and at a total pressure of approximately 1 atm [45]. The CO2 capture behavior at moderate temperature and pressure using the prepared porous nano-CaO adsorbent is a feasible preprocess, which determined the intrinsic CO2 capture capacity under typical postcombustion operating conditions. Furthermore, these results are of critical importance for carbon capture sequestration (CCS) technologies using adsorption processes.

Cyclic CO2/nano-CaO adsorption-desorption

Cycling processes may be the cause of CaO-based adsorbent decay. Therefore, knowing the stability of the adsorbent is necessary for CCS technology implementation on a large scale. A high stability would result in the reduction of capital cost and energy. The qad of five cycles was 11.0 ± 0.1 meq/g; therefore, the adsorbent could be regenerated five times without any significant loss of CO2 adsorption performance. Moreover, the material remains stable because the loss of the original samples after the five cycles was <2 wt%. Based on future energy scenarios, it is obvious that this porous nano-CaO adsorbent is an efficient material for balancing the costs and engineering feasibilities for different CO2 capture processes. Previous experimental reports over multiple cycles of adsorption-desorption activity at different adsorption temperatures of CaO-based adsorbent showed a loss of carbonation activity with a high number of cycles [46].

Conclusions

The new porous nano-CaO adsorbent synthesized by chemical combustion/ball milling with improved structural, chemical, and textural characteristics was able to absorb CO2 with excellent CO2 adsorption capacity at moderate temperatures. According to the Langmuir isotherms, this parameter increased from 9.48 mmol/g to 12.54 mmol/g when the temperature increased from 298 K to 323 K and to 348 K at ambient pressure. This achieved CaCO3 formation on the porous nano-CaO adsorbent surface, which was accompanied by a very small fraction of physisorption. The adsorption of CO2 could be adjusted finely to the kinetic pseudo-second-order and the intraparticle diffusion models, and this confirmed that the chemical process is carried out in gradual steps. The Langmuir adsorption isotherm was more accurate in predicting CO2 adsorption on the porous CaO nanoadsorbent, which indicates monolayer coverage at the uniform surface adsorption sites characterized by chemisorption on a homogeneous material. The thermodynamic parameters showed that the positive sign of ΔH° in the studied temperature range indicated that the CO2 adsorption reaction was endothermic and the magnitude revealed both physisorption and chemisorption processes. The positive value of ΔS° showed that the randomness at the solid-gas interface increased during the CO2 adsorption process. The positive sign of the ΔG° values implies a non-spontaneous adsorption reaction, and the diminishing of the values with temperature suggests that the adsorption is favored at higher temperatures. The value of Ea is also in agreement with chemisorption as the adsorption mechanism. Five cycles of regeneration of the porous nano-CaO adsorbent did not cause any significant loss of performance or properties. Therefore, the nano-CaO adsorbent is a potential and efficient material to be applied for CO2 capture in large-scale postcombustion CO2 capture technologies. The obtained kinetic, equilibrium, and thermodynamic results provide valuable information for the design of improved adsorption schemes for CO2 capture technologies at moderate temperatures using a new high-performance porous CaO nanoadsorbent.

eISSN:
2083-134X
Langue:
Anglais