Next Article in Journal
Isolation of Bioactive Compounds (Carotenoids, Tocopherols, and Tocotrienols) from Calendula Officinalis L., and Their Interaction with Proteins and Oils in Nanoemulsion Formulation
Previous Article in Journal
Identification of Cellular Isoschaftoside-Mediated Anti-Senescence Mechanism in RAC2 and LINC00294
Previous Article in Special Issue
The Use of Various Types of Waste Paper for the Removal of Anionic and Cationic Dyes from Aqueous Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Carbon Dioxide Adsorption over Activated Biocarbons Derived from Lemon Peel

by
Karolina Kiełbasa
1,
Joanna Siemak
1,
Joanna Sreńscek-Nazzal
1,
Bestani Benaouda
2,
Banasri Roy
3,* and
Beata Michalkiewicz
1,*
1
Department of Catalytic and Sorbent Materials Engineering, Faculty of Chemical Technology and Engineering, West Pomeranian University of Technology in Szczecin, Pulaskiego 10, 70–322 Szczecin, Poland
2
Laboratory of Structure, Development, and Application of Molecular Materials (SEA2M), Faculty of Sciences and Technology, Abdelahmid Ibn Badis University of Mostaganem, P.O. Box 227, Mostaganem 27000, Algeria
3
Department of Chemical Engineering, Birla Institute of Technology and Science (BITS), Pilani 333031, Rajasthan, India
*
Authors to whom correspondence should be addressed.
Molecules 2024, 29(17), 4183; https://doi.org/10.3390/molecules29174183
Submission received: 20 August 2024 / Revised: 29 August 2024 / Accepted: 2 September 2024 / Published: 4 September 2024
(This article belongs to the Special Issue Advances in Innovative Adsorbents)

Abstract

:
The rising concentration of CO2 in the atmosphere is approaching critical levels, posing a significant threat to life on Earth. Porous carbons derived from biobased materials, particularly waste byproducts, offer a viable solution for selective CO2 adsorption from large-scale industrial sources, potentially mitigating atmospheric CO2 emissions. In this study, we developed highly porous carbons from lemon peel waste through a two-step process, consisting of temperature pretreatment (500 °C) followed by chemical activation by KOH at 850 °C. The largest specific surface area (2821 m2/g), total pore volume (1.39 cm3/g), and micropore volume (0.70 cm3/g) were obtained at the highest KOH-to-carbon ratio of 4. In contrast, the sample activated with a KOH-to-carbon ratio of 2 demonstrated the greatest micropore distribution. This activated biocarbon exhibited superior CO2 adsorption capacity, reaching 5.69 mmol/g at 0 °C and 100 kPa. The remarkable adsorption performance can be attributed to the significant volume of micropores with diameters smaller than 0.859 nm. The Radke–Prausnitz equation, traditionally employed to model the adsorption equilibrium of organic compounds from liquid solutions, has been shown to be equally applicable for describing the gas–solid adsorption equilibrium. Furthermore, equations describing the temperature dependence of the Radke–Prausnitz equation’s parameters have been developed.

1. Introduction

The augmentation of environmental pollution and the depletion of fossil fuel sources are two very serious threats for the existence of the biosphere as well as human civilization. Environmental pollution is deeply connected to the anthropogenic usage of fossil fuels. According to the BP Statistical Review of World Energy 2019, 68th edition data, the global consumption rate of fossil fuels is 13,685 million tons (Mt)/year, with an annual growth rate of 2.8% [1]. In 2020, the global amounts of methane, NOx, and carbon dioxide were reported to be 1849.9 ppb, 330.31 ppb, and 410 ppm, respectively [2,3].
According to the United Nations Environmental Report 2019 [4], the total amount of greenhouse gases in the atmosphere reached 55 gigatons (Gt) and global temperature increased by 3 °C. Hence, the removal of CO2 from the environment is crucial. Depending upon the principle of the capture process, there are different types of techniques to separate and capture CO2 from the flue gas stream: absorption by means of solvents, membranes, and cryogenics; adsorption by solid materials; capture by microalgae; the calcium looping cycle, etc. Among these, the adsorption of CO2, especially at the post-combustion stage, by suitable solid materials with a high surface area such as carbonaceous materials, zeolite, metal–organic frameworks, silica-based materials, etc. could be energetically favorable for industrial applications [5].
Not only as an adsorbent, carbonaceous materials (activated carbon, graphene oxide, CNT, etc.) have attracted great attention due to their wide range of applications such as in energy storage (supercapacitors, batteries, etc.), materials reinforcement, templates for hollow structures, capsules for magnetic nanoparticles, environmental pollution control (both in the gas phase and in solution), drug delivery, etc. [6,7,8,9,10,11].
The improvement in these applications strongly depends on the ability to control the pore size, bulk structure, crystallinity, porosity, and surface properties of the carbon materials. Activated carbons (ACs) are receiving considerable attention for their various impending advantages such as high wear resistance, high mechanical strength, good adsorption performance, high purity, low ash content, high pore volume, and controllable pore size distribution [12,13].
Naturally abundant lignocellulosic biomass wastes (rice hull, corn husk, hazel nutshell, fruit peels, etc.) could be good precursor materials for the synthesis of activated carbon [14,15,16,17,18,19,20,21,22,23,24,25].
These biomass waste resources generate rapidly, are easy to access, and are environmentally benign in nature. Statistically, the European Union produces 7 million tons of orange and other easy-peel citrus fruits [26]. Over the years 2000–2021, the production of lemons and limes increased from 11.38 to 21.55 million tons. However, in 2022, it amounted to 21.53 million tons, which represents a slight decrease compared to the previous year’s production volume [27]. Hence, the production rate of the citrus peel, considered to be roughly 50% of the wet fruit mass, is around 3.5 million tons per year [28].
Several research articles reported the successful production of ACs from lemon peels with high surface areas and porosities.
Divahar et al. [29] developed activated carbon synthesized from lemon peels by the thermal activation method and investigated its lead ion removal efficiency from E-waste bioleachate. The activated carbon had a surface area of 603.7 m2/g, a total pore volume of 0.649 cm3/g, and an average pore diameter of 0.1 nm.
Weldekidan et al. [30] synthesized highly porous activated carbons from lemon peel waste using a two-step activation process involving hydrothermal carbonization followed by chemical activation with KOH and KOH + aluminum sulfate at 800 °C. The sample activated with KOH + aluminum had the highest BET surface area (2143 m2/g) and total pore volume of 1.3 cc/g. While the sample with only KOH activation had a lower surface area (1113 m2/g), it had better CO2 adsorption capacity, ranging up to 4.5 mmol/g, because of its well-developed pore structures and a high proportion (90%) of micropore distributions.
Yusop et al. [31] synthesized activated carbon from lemon peels by pyrolysis followed by the CO2 activation method and Cu salt modification. These materials were used for studying the adsorption of amoxicillin (AMOX) in wastewater. The Cu-modified activated carbon exhibited the maximum BET surface area, pore volume, and pore diameter of 1018 m2/g, 0.5377 cm3/g, and 2.78 nm, respectively.
De Rose et al. [32] synthesized activated carbon from lemon peels by pyrolysis and the physical activation (in a CO2 atmosphere) method, and they studied their hydrogen adsorption characteristics. The maximum surface area and total pore volume were found to be 500 m2/g and 0.208 cm3/g, respectively.
Li et al. [33] reported on the synthesis of magnetic biochar composite from lemon peel residue by a one-step hydrothermal method, and it was used for the adsorptive capture of methylene blue in an aqueous solution.
Ramutshatsha-Makhwedzha et al. [34] synthesized activated carbon from the 1:1 mixture of waste orange and lemon peel using the phosphoric acid (H3PO4) activation process and applied it to the removal of Cd2+, Pb2+, and Fe2+ ions from acid mine drainage. The activated carbon had a BET surface area and pore size of 168 m2/g and 5.18 nm, respectively.
Ramutshatsha-Makhwedzha et al. [35]) synthesized activated carbon from mixed orange and lemon peels using phosphoric acid and used it for the removal of methyl orange and methylene dyes from wastewater. The BET surface area, pore volume, and pore size of the activated carbon were measured to be 168.29 m2/g, 0.269 m3/g, and 5.18 nm, respectively.
Praipipat et al. [36] reported on the application of lemon peel beads-doped iron (III) oxide-hydroxide (LBF) and lemon peel beads-doped zinc oxide (LBZ) for the removal of blue 4 dye from water.
Surya and Michael [37] demonstrated the preparation of activated carbon from lemon peels by carbonization at 400 °C followed by chemical activation employing KOH, and the activated carbon was utilized for the fabrication of the electrode of an electrochemical double-layer capacitor.
Naser et al. [38] reported on the synthesis of activated lemon peels using concentrated sulfuric acid and its application for the removal of Pb (II) ions from water.
Su et al. [39] synthesized carbon quantum dots (CQDs) from dried lemon peels using a hydrothermal treatment at 200 °C for a period of 6 h under normal water vapor pressure. The obtained CQDs were utilized for the development of a carmine determination probe sensor.
Sharifzade et al. [40] studied the synthesis of activated carbon from lemon citrus peels using phosphoric acid at 500 °C for the removal of xanthene dyes, erythrosine B (EB), and rhodamine B (RB) from an aqueous binary dye solution.
Mohammadi et al. [41] prepared activated carbon from lemon peels by the chemical activation method using H3PO4. The BET surface area was revealed to be 1158 m2/g.
According to our knowledge, CO2 adsorption was investigated over activated carbons derived from lemon peels only once. The lemon peels were converted to activated carbons via hydrothermal carbonization coupled with chemical activation using KOH at 800 °C for CO2 adsorption [30]. The KOH–carbon precursor ratio was equal to 2. The specific surface area of 1113 m2/g and a CO2 adsorption of 4.5 mmol/g at 0 °C and 1 bar were achieved.
Given the very large supply of lemon peel waste worldwide and the limited possibilities for its utilization, we decided to produce activated carbon from this precursor. Moreover, considering the unresolved problem of finding an effective CO2 sorbent, we aimed to develop a synthesis of active biocarbon targeted at CO2 adsorption. Instead of a hydrothermal pretreatment, we applied a high-temperature pretreatment (500 °C) and demonstrated that our method guaranteed the production of a sorbent with higher CO2 adsorption than that obtained by Weldekidan et al. [30], who, like us, also used lemon peels as the carbon precursor and KOH as the activating agent.
The absolute novelty we present here is the application of the Radke–Prausnitz equation to the gas–solid adsorption equilibrium and the demonstration that the parameters of this equation depend on temperature. We also propose forms of these dependencies.

2. Results

2.1. XRD Results

The XRD spectra (Figure 1) depict the semicrystalline nature of the activated biocarbons derived from lemon peels under different treatment conditions.
A brief explanation of the activated biocarbons and char derived from the lemon peels under different treatment conditions are presented in Table 1. The detailed explanations are presented in Section 3 of Materials and Methods.
Two prominent peaks at ~23.8° and ~43.6° tally to the (002) and (100) planes (ICSD #76767) of graphitic carbon, respectively. The appearance of these peaks indicates that the activated biocarbons produced from lemon peels were primarily amorphous in nature. After treatment with KOH, the peak at ~23.8° disappeared. This indicates that the individual layers are arranged in a disordered manner relative to each other. However, the signal at ~43.6° became more pronounced. It can be assumed that the average size of the layers increased.

2.2. Textural Properties

Figure 2 reveals the N2 adsorption–desorption isotherms at −196 °C for the activated biocarbons from lemon peels obtained under different treatment conditions. LP500 showed a very low specific surface area (Table 1) of 0.238 m2/g with a total pore volume of 0.003 cm3/g. For the pretreated KOH activated biocarbons, after p/p0 = 0.3, the adsorption–desorption profiles extended to a flat plateau with the increase in p/p0 values generating a classic type I adsorption isotherm, indicating the presence of a very high amount of micropores in the carbon structure. The only KOH-activated sample displayed a slight hysteresis loop between p/p0 0.4 and 0.7, and for the other KOH-treated samples, the adsorption data overlapped with the desorption data.
Other samples showed high N2 adsorption, indicating a high BET surface area; however, the biochar pretreatment of the lemon peels significantly enhanced the surface area and pore volume of the activated biocarbon. Lemon peels with only KOH (KOH–lemon peel mass ratio = 1:1) activation (LP_1) had a surface area and pore volume of 1087 m2/g and 0.719 cm3/g, respectively (Table 2). However, for the sample pretreated at 500 °C followed by KOH activation of same ratio, an almost 80% enhancement of surface area (1835 m2/g) and around a 27% enhancement in total pore volume (0.91 cm3/g) were observed. In the pretreated samples, these values increased with the KOH–lemon peel mass ratio, and the largest surface area (2821 m2/g) and pore volume (1.39 cm3/g) were achieved for KOH–lemon peel = 4:1.
When comparing the textural properties of the activated biocarbons from lemon peels activated by KOH produced by us with activated biocarbons obtained from the same carbon source and activated with the same precursor [30], it must be noted that regardless of the amount of KOH used, our materials exhibited higher specific surface area, pore volume, and micropore volume. The key parameter, namely carbon adsorption at 0 °C under 1 bar pressure, was also higher. Most likely, the pretreatment method, specifically the temperature, plays a crucial role.
Figure 3 shows the pore size distributions of the samples. In the micropore range for materials activated with KOH, the highest pore volume was observed in the ranges of 0.4–0.6 nm and 0.8–0.9 nm. The material LP500_2 had the largest pore volume in the 0.4–0.6 nm range. For the only KOH-activated sample, pores were mostly in the 2–4 and 5–8 nm range, whereas for the pretreated and KOH-activated samples, a large portion of the pores were less than 1 nm and in the 2–4 nm size range.

2.3. SEM Results

SEM pictures of activated biocarbons and char derived from lemon peels are presented in Figure 4. In the SEM image of the LP500 sample, which was not activated in any way and is simply a carbonized lemon peel, the formation of a non-porous char and a smooth, unwrinkled surface were clearly visible. The SEM images of char that was activated using various amounts of KOH showed increasingly pronounced surface wrinkling as the amount of activator used increased. In the case of using a larger amount of KOH (LP500_4), the surface of the material was composed of thin carbon sheets randomly connected. The material obtained without the preliminary thermal treatment of lemon peels appeared completely different. Treating the carbon precursor with an activating agent and subjecting it to carbonization (LP_1) resulted in a material with a completely different morphology than described above, namely a slightly wrinkled surface characterized by numerous macropores. The SEM images confirm that all chemically activated materials exhibited high porosity, with LP500_4 being the most porous. This is consistent with the N2 sorption results presented in Table 1.

2.4. CO2 Adsorption

Figure 5 shows the CO2 adsorption profiles of the activated biocarbons studied at ambient pressure and 0 °C. The maximum CO2 uptake (5.69 mmol/g) happened for the pretreated and KOH-activated biocarbon sample with KOH–lemon peel mass ratio = 2:1.
This value is quite high, considering that the synthesis of these activated carbons is relatively simple and inexpensive. If we compare the CO2 adsorption values under the same conditions for MOF materials, i.e., 1.67 mmol/g [42] and 3.07 mmol/g [43], whose production is far more complex, it further emphasizes the success of activated biocarbons from lemon peels.
It is important to note here that the maximum surface area and total pore volume were obtained for the pretreated and KOH-activated biocarbon sample with KOH–lemon peel mass ratio = 4:1. However, the high proportion of the micropore distributions (70%) in the pretreated and KOH-activated biocarbon sample with KOH–lemon peel mass ratio = 2:1 (Table 1), along with the narrower pore size distributions (Figure 3), resulted in better CO2 adsorption performance of the sample.
To better understand the influence of textural parameters on CO2 sorption at a temperature of 0 °C and a pressure of 100 kPa, we examined the data presented in Figure 6. Relationships between CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa, and specific surface area (SSA), total pore volume (Vtot), and micropore volume (Vmicro) show CO2 adsorption as a function of specific surface area, total pore volume, and micropore volume.
Casco et al. [44] demonstrated a good correlation between CO2 adsorption and surface area, total pore volume, and micropore volume. The R2 value was higher than 0.9 for all these textural parameters. Nevertheless, the best correlation (R2 = 0.993) was observed for specific surface area. Our studies did not confirm this. The highest, but still unsatisfactory, R2 (0.896) value was obtained for micropore volume. The reason may be that the aforementioned authors conducted adsorption studies at a temperature of 25 °C and a pressure of 4.5 MPa. It was shown [45] that, depending on the pressure and temperature conditions, different textural parameters become significant for efficient CO2 adsorption.
Considering that in our studies, the R2 value suggests that micropores may have some significance for CO2 adsorption. We decided to focus on a detailed analysis of the influence of micropore volume with precisely defined diameters.
The relationship between CO2 adsorption at 1 bar and 0 °C and the volume of specific pores smaller than a certain diameter was analyzed (Figure 7). The higher coefficient of determination was observed for the volume of pores smaller than 0.859 nm, resulting in an R2 of 0.999. The relationships between CO2 adsorption and pore diameter for the three highest R2 values are presented in Figure 8. The CO2 adsorption capacities examined by other authors were primarily attributed to the highly developed microporosity within the range below 0.8 [46,47] and 0.82 [48]. Rouzitalab et al. [45] analyzed over 200 carbon materials and found an apparent discrepancy with the concept that a higher surface area would lead to greater CO2 adsorption. In the same comprehensive review, the importance of pore size was highlighted, indicating that the highest CO2 adsorption was achieved for materials with a high pore volume in the diameter range of less than 0.8–0.9 nm. Our research, although presented for only four carbon materials, confirms the results reported by other authors.
The experimental results displayed in Figure 9 were analyzed using the Freundlich, Langmuir, Toth, Sips, UNILAN, and Radke–Prausnitz equations. Five error functions (SSE, HYBRID, ARE, MPSD, and SAE) were applied to find the best-fit model. The Freundlich, Langmuir, Toth, Sips, UNILAN, and error functions were described in detail by Siemak and Michalkiewicz [49]. Since several different equations referred to as Radke–Prausnitz can be found in the literature, it is important to go back to the origins and clarify that the Radke–Prausnitz model was first described by Radke and Prausnitz [50] for the adsorption of organic solutes from a dilute aqueous solution onto activated carbon, and the equation was originally expressed as (1):
1 n i c = 1 a · c i + 1 b · c i β
where
n¡c is the amount of solute adsorbed per unit mass of adsorbent, expressed in millimoles per gram;
ci is the solute concentration, expressed in moles per liter;
ß is the Radke–Prausnitz parameter restricted to be less than unity.
Authors that later utilized the Radke–Prausnitz equation to other solutions of organic compounds rearranged it to (2) [51,52,53]:
1 q e = 1 a R P · C e + 1 b R P · C e β
By rearranging Equation (2), Equation (3) can be obtained, which represents the nonlinear expression of the Radke–Prausnitz equation [54,55,56].
q = a R P · b R P · C e β a R P + b R P · C e β 1
The common Radke–Prausnitz model used for adsorption from liquid solutions has been adapted by us for gas adsorption. Based on Equation (3), Equation (4) can be derived as
q = a · b · p n a + b · p n 1
where:
q is CO2 equilibrium adsorption at p;
p is the equilibrium pressure;
a [mmol/g], b [kPa−1], n—Radke-Rrausnitz model constants
Solver, a tool available in Excel, was utilized for finding the parameters and analyzing error functions. The values of the error functions for all the equations listed above are presented in Table 3. The lowest error function values were obtained for the Radke–Prausnitz equation at all temperatures. Among all the models, the Radke–Prausnitz model provided the best fit to the experimental data. Table 4 demonstrate the parameters of the Radke–Prausnitz model and errors of the model.
To effectively describe adsorption equilibrium data across different temperatures, it is essential to have an isotherm equation that accounts for temperature dependence. The temperature dependences of parameters in equations describing adsorption equilibria, such as Freundlich, Langmuir, Sips, Toth, and UNILAN, can be found in the literature [57]. However, such dependencies for the Radke–Prausnitz equation have not yet been described. We have proposed temperature dependencies (Equations (5)–(7)) of the a, b, and n parameters from Equation (4) based on the similarity of the Sips and Toth equations to the Radke–Prausnitz equation:
a = a 0 · e x p χ 1 T T 0
b = b 0 e x p Q R T 0 T 0 T 1
n = n 0 + α 1 T 0 T
In Equations (5)–(7), a0, χ, Q, b0, n0, and α are the constants. R represents the ideal gas constant. T0 is the reference temperature. To verify whether the proposed equations can be applied to the Radke–Prausnitz parameters, plots of the natural logarithm of a versus temperature, the natural logarithm of b versus the inverse of temperature, and n versus the inverse of temperature were created and are shown in Figure 10. A temperature of 0 °C was chosen as the reference temperature.
On the basis of the plots presented in Figure 10, the parameters a0, α, χ, Q, b0, and n0 were calculated and presented in Table 5.
The Clausius–Clapeyron equation was employed to calculate the isosteric heat of adsorption:
Q i s o = R ln p 1 T θ
where θ is the degree of surface coverage.
After integration of Equation (8), the linear function (9) is obtained:
ln p θ = Q i s o R 1 T + C
On the basis of the Radke–Prausnitz Equation (4) and the parameters listed in Table 2, the adsorption isosteres (9), i.e., the natural logarithm of pressure values plotted against the inverse of temperature for different fixed surface coverages, were plotted in Figure 11.
Using the slopes of the resulting straight lines, the isosteric heat of adsorption was determined. The isosteric heat of adsorption at different levels of surface coverage is illustrated in Figure 12.
The isosteric heat of adsorption (Figure 12) was determined through adsorption experiments conducted at the five different temperatures presented in Figure 9. As CO2 loading increased, the isosteric heat of adsorption declined. This occurs because the strongest adsorption sites become saturated, leading CO2 to be adsorbed onto weaker sites. The calculated values of the isosteric heat of adsorption ranged from 25 to 30 kJ/mol. Such low values indicate that CO2 adsorption on activated biocarbon LP500_2 is of a physical nature (physisorption). Suitable adsorbents for CO2 capture must, among other things, meet the requirement of gentle regeneration conditions [58]. The ease of regenerating the adsorbent is a crucial factor when selecting materials for CO2 separation. A low heat of adsorption ensures that this condition is met. This value of heat adsorption indicates the promise of the activated biocarbon LP500_2 to be used in low-pressure carbon dioxide adsorption applications.

3. Materials and Methods

3.1. Starting Materials

For the synthesis of activated biocarbon powder from lemon peels, the following chemical reagents were used: potassium hydroxide (Chempur, pure p.a.) and 35–38% hydrochloric acid (Chempur, pure p.a.).

3.2. Preparation of Activated Biocarbons from Lemon Peels

Lemons (Citrus limon) were collected from a local market, peels were separated manually, and cut into small pieces of ~0.5 cm × 0.5 cm × 0.5 cm. These were air dried for 48 h and then oven dried at 60 °C for 24 h. The dried peels were crushed in a coffee grinder and strained to collect the powder of ~100–200 µm. The powder was heat treated in two ways. In the first method, it was homogeneously mixed with saturated KOH (KOH–lemon peel mass ratio = 1:1) solution and kept at room temperature for 3 h followed by oven drying at 200 °C for 19 h. The dried mass was carbonized at 850 °C for 1 h under N2 flow of 14.4 dm3/h. The black mass was collected, transferred to a Buchner funnel, and washed with distilled water until the pH became neutral. The wet black powder was transferred to a 500 cm3 beaker containing 100 cm3 1M HCl solution, and the excess solution was evaporated on a hot plate. Again, it was transferred to a Buchner funnel and washed with distilled water until the pH became neutral. Finally, the wet black mass was transferred to a 500 cm3 beaker with water, and the excess water was evaporated and dried at 200 °C overnight. The sample was labeled as LP_1. Sample LP500 was prepared similarly but without the addition of KOH.
In the second method, the dried lemon peel powder was first carbonized at 500 °C for 1 h under N2 flow of 14.4 cm3/h and then impregnated with KOH (KOH–lemon peel mass ratio varied as 1:1, 1:2, and 1:4) and chemically activated in the same way as in the first method. The samples were labeled as LP500_1, LP500_2, and LP500_4. The last symbol refers to the amount of KOH.

3.3. Characterization of the Activated Biocarbon

An X-ray diffractometer (X’Pert–PRO, Panalytical, Almelo, The Netherlands, 2012) was used to study the phase of the sample within the range of 2θ = 10–100°. The scanning rate was 0.022349 °/s. The textural properties (specific surface area, pore volume, porosity) of the samples were investigated by a N2 adsorption–desorption analyzer (Quadrasorb evo™ Gas Sorption analyzer, Anton Paar, St Albans, UK; previously Quantachrome Instruments, Boynton Beach, FL, USA, 2014). The specific surface area was calculated using the Brunauer–Emmett–Teller (BET) equation from P/P0 = 0.05 to 0.2. The total pore volume was calculated according to the N2 adsorbed at P/P0 = 1.
The micropore (1.3–2 nm) volume (Vmicro) was calculated by the DFT method. Total CO2 adsorption was determined at 298 K and 1 bar. A field emission scanning electron microscope (SU8020 Ultra-High Resolution Field Emission Scanning Electron Microscope, Hitachi Ltd., Tokyo, Japan, 2012) using secondary and backscattered electron detectors was used to examine the morphology and surface features of the obtained activated biocarbon.

4. Conclusions

Lemon peel waste was transformed into highly porous and especially microporous biocarbons through temperature pretreatment (500 °C) followed by chemical activation with KOH at 850 °C, specifically for CO2 adsorption. All chemically activated samples had a well-developed micropore structure. The highest specific surface area (2821 m2/g), total pore volume (1.39 cm3/g), and micropore volume (0.70 cm3/g) were achieved for the highest KOH-to-carbon source ratio (4).
The sample activated with a KOH-to-carbon source ratio of 2 exhibited the highest proportion of micropore distributions. This particular activated biocarbon was the best CO2 sorbent (5.69 mmol/g at 0 °C and 100 kPa). Such a high adsorption was achieved due to the large volume of micropores smaller than 0.859 nm.
It has been demonstrated that the Radke–Prausnitz equation, previously used to describe the adsorption equilibrium of organic compounds from solutions, can also be applied to describe the gas–solid adsorption equilibrium. Additionally, equations describing the dependence of the Radke–Prausnitz equation parameters on temperature were proposed.

Author Contributions

Conceptualization, B.R. and B.M.; methodology, B.R. and B.M.; formal analysis, J.S. and B.B.; investigation, B.R., K.K. and J.S.-N.; data curation, B.R. and J.S.; writing—original draft preparation, B.R., B.M. and J.S.; writing—review and editing, B.R. and B.M.; visualization, J.S.; supervision, B.R. and B.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original data presented in this study are openly available in Mendeley Data at https://data.mendeley.com/datasets/94jg7jkt6n/1 (accessed on 8 August 2024).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dale, S. BP Statistical Review of World Energy, 68th ed.; Plance of Publication: London, UK, 2019. [Google Scholar]
  2. European Environment Agency. Trends in Atmospheric Concentrations of CO2 (Ppm), CH4 (Ppb) and N2O (Ppb), between 1800 and 2017. Available online: https://www.eea.europa.eu/data-and-maps/daviz/atmospheric-concentration-of-carbon-dioxide-5#tab-chart_5_filters=%7B%22rowFilters%22%3A%7B%7D%3B%22columnFilters%22%3A%7B%22pre_config_polutant%22%3A%5B%22CH4(ppb)%22%5D%7D%7D (accessed on 1 August 2024).
  3. Page, M. Le New Scientist. 2020. Carbon Dioxide Levels Will Soar Past the 410 Ppm Milestone in 2019. Available online: https://www.newscientist.com/article/2191881-carbon-dioxide-levels-will-soar-past-the-410-ppm-milestone-in-2019/ (accessed on 1 August 2024).
  4. UNEP, UN Environment Programme, Cut Global Emissions by 7.6 Percent Every Year for next Decade to Meet 1.5 °C Paris Target–Unreport. Available online: https://www.unep.org/news-and-stories/press-release/cut-global-emissions-76-percent-every-year-next-decade-meet-15degc (accessed on 1 August 2024).
  5. Raganati, F.; Miccio, F.; Ammendola, P. Adsorption of Carbon Dioxide for Post-Combustion Capture: A Review. Energy Fuels 2021, 35, 12845–12868. [Google Scholar] [CrossRef]
  6. Chen, J.; Xia, N.; Zhou, T.; Tan, S.; Jiang, F.; Yuan, D. Mesoporous Carbon Spheres: Synthesis, Characterization and Supercapacitance. Int. J. Electrochem. Sci. 2009, 4, 1063–1073. [Google Scholar] [CrossRef]
  7. Jin, Y.Z.; Kim, Y.J.; Gao, C.; Zhu, Y.Q.; Huczko, A.; Endo, M.; Kroto, H.W. High Temperature Annealing Effects on Carbon Spheres and Their Applications as Anode Materials in Li-Ion Secondary Battery. Carbon 2006, 44, 724–729. [Google Scholar] [CrossRef]
  8. Sun, X.; Liu, J.; Li, Y. Use of Carbonaceous Polysaccharide Microspheres as Templates for Fabricating Metal Oxide Hollow Spheres. Chem. Eur. J. 2006, 12, 2039–2047. [Google Scholar] [CrossRef]
  9. Wang, Z.; Xiao, P.; He, N. Synthesis and Characteristics of Carbon Encapsulated Magnetic Nanoparticles Produced by a Hydrothermal Reaction. Carbon 2006, 44, 3277–3284. [Google Scholar] [CrossRef]
  10. Song, X.; Gunawan, P.; Jiang, R.; Leong, S.S.J.; Wang, K.; Xu, R. Surface Activated Carbon Nanospheres for Fast Adsorption of Silver Ions from Aqueous Solutions. J. Hazard. Mater. 2011, 194, 162–168. [Google Scholar] [CrossRef] [PubMed]
  11. Wen, Z.; Wang, Q.; Zhang, Q.; Li, J. Hollow Carbon Spheres with Wide Size Distribution as Anode Catalyst Support for Direct Methanol Fuel Cells. Electrochem. Commun. 2007, 9, 1867–1872. [Google Scholar] [CrossRef]
  12. Wang, Q.; Ling, X.; Zhang, R.; Liu, C.; Liu, X.; Qiao, W.; Zhan, L.; Ling, L. Preparation of Polystyrene-Based Activated Carbon Spheres and Their Adsorption of Dibenzothiophene. New Carbon Mater. 2009, 24, 55–60. [Google Scholar] [CrossRef]
  13. Liu, Z.; Ling, L.; Qiao, W.; Liu, L. Effect of Hydrogen on the Mesopore Development of Pitch-Based Spherical Activated Carbon Containing Iron during Activation by Steam. Carbon 1999, 37, 2063–2066. [Google Scholar] [CrossRef]
  14. Serafin, J.; Kiełbasa, K.; Michalkiewicz, B. The New Tailored Nanoporous Carbons from the Common Polypody (Polypodium Vulgare): The Role of Textural Properties for Enhanced CO2 Adsorption. Chem. Eng. J. 2022, 429, 131751. [Google Scholar] [CrossRef]
  15. Serafin, J.; Dziejarski, B.; Vendrell, X.; Kiełbasa, K.; Michalkiewicz, B. Biomass Waste Fern Leaves as a Material for a Sustainable Method of Activated Carbon Production for CO2 Capture. Biomass Bioenergy 2023, 175, 106880. [Google Scholar] [CrossRef]
  16. Elaiyappillai, E.; Srinivasan, R.; Johnbosco, Y.; Devakumar, P.; Murugesan, K.; Kesavan, K.; Johnson, P.M. Low Cost Activated Carbon Derived from Cucumis Melo Fruit Peel for Electrochemical Supercapacitor Application. Appl. Surf. Sci. 2019, 486, 527–538. [Google Scholar] [CrossRef]
  17. Yogin Soodesh, C.; Seriyala, A.K.; Navjot; Chattopadhyay, P.; Rozhkova, N.; Michalkiewicz, B.; Chatterjee, S.; Roy, B. Carbonaceous Catalysts (Biochar and Activated Carbon) from Agricultural Residues and Their Application in Production of Biodiesel: A Review. Chem. Eng. Res. Des. 2024, 203, 759–788. [Google Scholar] [CrossRef]
  18. Akula, S.; Varathan, P.; Menon, R.S.; Sahu, A.K. Rationally Constructing Nitrogen–Fluorine Heteroatoms on Porous Carbon Derived from Pomegranate Fruit Peel Waste towards an Efficient Oxygen Reduction Catalyst for Polymer Electrolyte Membrane Fuel Cells. Sustain. Energy Fuels 2021, 5, 886–899. [Google Scholar] [CrossRef]
  19. Gandla, D.; Wu, X.; Zhang, F.; Wu, C.; Tan, D.Q. High-Performance and High-Voltage Supercapacitors Based on N-Doped Mesoporous Activated Carbon Derived from Dragon Fruit Peels. ACS Omega 2021, 6, 7615–7625. [Google Scholar] [CrossRef]
  20. Jothi Ramalingam, R.; Sivachidambaram, M.; Vijaya, J.J.; Al-Lohedan, H.A.; Muthumareeswaran, M.R. Synthesis of Porous Activated Carbon Powder Formation from Fruit Peel and Cow Dung Waste for Modified Electrode Fabrication and Application. Biomass Bioenergy 2020, 142, 105800. [Google Scholar] [CrossRef]
  21. Jawad, A.H.; Saud Abdulhameed, A.; Wilson, L.D.; Syed-Hassan, S.S.A.; ALOthman, Z.A.; Rizwan Khan, M. High Surface Area and Mesoporous Activated Carbon from KOH-Activated Dragon Fruit Peels for Methylene Blue Dye Adsorption: Optimization and Mechanism Study. Chinese J. Chem. Eng. 2021, 32, 281–290. [Google Scholar] [CrossRef]
  22. Ahmad, M.A.; Eusoff, M.A.; Oladoye, P.O.; Adegoke, K.A.; Bello, O.S. Statistical Optimization of Remazol Brilliant Blue R Dye Adsorption onto Activated Carbon Prepared from Pomegranate Fruit Peel. Chem. Data Collect. 2020, 28, 100426. [Google Scholar] [CrossRef]
  23. Sahu, S.; Pahi, S.; Sahu, J.K.; Sahu, U.K.; Patel, R.K. Kendu (Diospyros Melanoxylon Roxb) Fruit Peel Activated Carbon—An Efficient Bioadsorbent for Methylene Blue Dye: Equilibrium, Kinetic, and Thermodynamic Study. Environ. Sci. Pollut. Res. 2020, 27, 22579–22592. [Google Scholar] [CrossRef]
  24. Arie, A.A.; Kristianto, H.; Cengiz, E.C.; Demir-Cakan, R. Activated Porous Carbons Originated from the Indonesian Snake Skin Fruit Peel as Cathode Components for Lithium Sulfur Battery. Ionics 2019, 25, 2121–2129. [Google Scholar] [CrossRef]
  25. Murugesan, K.; Tareke, K.; Gezehegn, M.; Kebede, M.; Yazie, A.; Diyana, G. Rapid Development of Activated Carbon and ZnO Nanoparticles via Green Waste Conversion Using Avocado Fruit Peel Powder and Its High Performance Efficiency in Aqueous Dye Removal Application. J. Inorg. Organomet. Polym. Mater. 2019, 29, 1368–1374. [Google Scholar] [CrossRef]
  26. Miserius, M.; Behr, H.-C. Fruitlogistica 2021 European Statistics Handbook; Messe Berlin GmbH: Berlin, Germany, 2021. [Google Scholar]
  27. Consumer Goods & FMCG› Food & Nutrition. Available online: https://www.statista.com/statistics/577445/world-lemon-and-lime-production/ (accessed on 1 August 2024).
  28. John, I.; Muthukumar, K.; Arunagiri, A. A Review on the Potential of Citrus Waste for D-Limonene, Pectin, and Bioethanol Production. Int. J. Green Energy 2017, 14, 599–612. [Google Scholar] [CrossRef]
  29. Divahar, R.; Meenambal, T.; Senophiyah Mary, J.; Aravind Raj, P.S.; Sangeetha, S.P.; Athavan Alias Anand, S. Lemon Peel Activated Carbon: A Sustainable Solution for Lead Ion Removal from E-Waste Bioleachate. Sustain. Chem. Environ. 2024, 6, 100094. [Google Scholar] [CrossRef]
  30. Weldekidan, H.; Patel, H.; Mohanty, A.; Misra, M. Synthesis of Porous and Activated Carbon from Lemon Peel Waste for CO2 Adsorption. Carbon Capture Sci. Technol. 2024, 10, 100149. [Google Scholar] [CrossRef]
  31. Mohamad Yusop, M.F.; Abdullah, A.Z.; Ahmad, M.A. Amoxicillin Adsorption from Aqueous Solution by Cu(II) Modified Lemon Peel Based Activated Carbon: Mass Transfer Simulation, Surface Area Prediction and F-Test on Isotherm and Kinetic Models. Powder Technol. 2024, 438, 119589. [Google Scholar] [CrossRef]
  32. De Rose, E.; Bartucci, S.; Poselle Bonaventura, C.; Conte, G.; Agostino, R.G.; Policicchio, A. Effects of Activation Temperature and Time on Porosity Features of Activated Carbons Derived from Lemon Peel and Preliminary Hydrogen Adsorption Tests. Colloids Surfaces A Physicochem. Eng. Asp. 2023, 672, 131727. [Google Scholar] [CrossRef]
  33. Li, Z.; Zhou, M.; Liu, N.; Zhang, F.; An, K.; Xiong, X.; Fan, S.; Sun, Q.; Le, T. Engineered Biochar Derived from Lemon Peel Waste for Highly Efficient Removal of Organic Pollutants from Water. Arab. J. Chem. 2023, 16, 105158. [Google Scholar] [CrossRef]
  34. Ramutshatsha-Makhwedzha, D.; Mbaya, T.; Mavhungu, A.; Mavhunga, M.L.; Mbaya, R. Adsorptive Removal of Cd2+, Pb2+, and Fe2+ from Acid Mine Drainage Using a Mixture of Waste Orange and Lemon Activated Carbon (WOLAC): Equilibrium Study. J. Iran. Chem. Soc. 2023, 20, 1119–1133. [Google Scholar] [CrossRef]
  35. Ramutshatsha-Makhwedzha, D.; Mavhungu, A.; Moropeng, M.L.; Mbaya, R. Activated Carbon Derived from Waste Orange and Lemon Peels for the Adsorption of Methyl Orange and Methylene Blue Dyes from Wastewater. Heliyon 2022, 8, e09930. [Google Scholar] [CrossRef]
  36. Praipipat, P.; Ngamsurach, P.; Prasongdee, V. Comparative Reactive Blue 4 Dye Removal by Lemon Peel Bead Doping with Iron(III) Oxide-Hydroxide and Zinc Oxide. ACS Omega 2022, 7, 41744–41758. [Google Scholar] [CrossRef]
  37. Surya, K.; Michael, M.S. Hierarchical Porous Activated Carbon Prepared from Biowaste of Lemon Peel for Electrochemical Double Layer Capacitors. Biomass Bioenergy 2021, 152, 106175. [Google Scholar] [CrossRef]
  38. Naser, S.M.; Ali, S.A.; Alkizwini, R.S.; Alshamali, M.; Alquzweeni, S.S.; Abdellatif, M.; Amoako-Attah, J.; AlKhayyat, A. Agro-Based Carbon for Lead Removal from Solutions. IOP Conf. Ser. Mater. Sci. Eng. 2021, 1058, 012019. [Google Scholar] [CrossRef]
  39. Su, A.; Wang, D.; Shu, X.; Zhong, Q.; Chen, Y.; Liu, J.; Wang, Y. Synthesis of Fluorescent Carbon Quantum Dots from Dried Lemon Peel for Determination of Carmine in Drinks. Chem. Res. Chin. Univ. 2018, 34, 164–168. [Google Scholar] [CrossRef]
  40. Sharifzade, G.; Asghari, A.; Rajabi, M. Highly Effective Adsorption of Xanthene Dyes (Rhodamine B and Erythrosine B) from Aqueous Solutions onto Lemon Citrus Peel Active Carbon: Characterization, Resolving Analysis, Optimization and Mechanistic Studies. RSC Adv. 2017, 7, 5362–5371. [Google Scholar] [CrossRef]
  41. Mohammadi, S.Z.; Karimi, M.A.; Yazdy, S.N.; Shamspur, T.; Hamidian, H. Removal of Pb(Ii) Ions and Malachite Green Dye From Wastewater Yy Activated Carbon Produced From Lemon Peel. Quim. Nova 2014. [Google Scholar] [CrossRef]
  42. Fu, X.-P.; Wang, Y.-L.; Zhang, X.-F.; Krishna, R.; He, C.-T.; Liu, Q.-Y.; Chen, B. Collaborative Pore Partition and Pore Surface Fluorination within a Metal–Organic Framework for High-Performance C2H2/CO2 Separation. Chem. Eng. J. 2022, 432, 134433. [Google Scholar] [CrossRef]
  43. Yang, S.-Q.; Krishna, R.; Chen, H.; Li, L.; Zhou, L.; An, Y.-F.; Zhang, F.-Y.; Zhang, Q.; Zhang, Y.-H.; Li, W.; et al. Immobilization of the Polar Group into an Ultramicroporous Metal–Organic Framework Enabling Benchmark Inverse Selective CO2/C2H2 Separation with Record C2H2 Production. J. Am. Chem. Soc. 2023, 145, 13901–13911. [Google Scholar] [CrossRef]
  44. Casco, M.E.; Martínez-Escandell, M.; Silvestre-Albero, J.; Rodríguez-Reinoso, F. Effect of the Porous Structure in Carbon Materials for CO2 Capture at Atmospheric and High-Pressure. Carbon 2014, 67, 230–235. [Google Scholar] [CrossRef]
  45. Rouzitalab, Z.; Maklavany, D.M.; Jafarinejad, S.; Rashidi, A. Lignocellulose-Based Adsorbents: A Spotlight Review of the Effective Parameters on Carbon Dioxide Capture Process. Chemosphere 2020, 246, 125756. [Google Scholar] [CrossRef]
  46. Wickramaratne, N.P.; Jaroniec, M. Importance of Small Micropores in CO2 Capture by Phenolic Resin-Based Activated Carbon Spheres. J. Mater. Chem. A 2013, 1, 112–116. [Google Scholar] [CrossRef]
  47. Presser, V.; McDonough, J.; Yeon, S.H.; Gogotsi, Y. Effect of Pore Size on Carbon Dioxide Sorption by Carbide Derived Carbon. Energy Environ. Sci. 2011, 4, 3059. [Google Scholar] [CrossRef]
  48. Chen, T.; Deng, S.; Wang, B.; Huang, J.; Wang, Y.; Yu, G. CO2 Adsorption on Crab Shell Derived Activated Carbons: Contribution of Micropores and Nitrogen-Containing Groups. RSC Adv. 2015, 5, 48323–48330. [Google Scholar] [CrossRef]
  49. Siemak, J.; Michalkiewicz, B. Adsorption Equilibrium of CO2 on Microporous Activated Carbon Produced from Avocado Stone Using H2SO4 as an Activating Agent. Sustainability 2023, 15, 16881. [Google Scholar] [CrossRef]
  50. Radke, C.J.; Prausnitz, J.M. Adsorption of Organic Solutes from Dilute Aqueous Solution of Activated Carbon. Ind. Eng. Chem. Fundam. 1972, 11, 445–451. [Google Scholar] [CrossRef]
  51. Farrier, D.S.; Hines, A.L.; Wang, S.W. Adsorption of Phenol and Benzoic Acid from Dilute Aqueous Solution onto a Macroreticular Resin. J. Colloid Interface Sci. 1979, 69, 233–237. [Google Scholar] [CrossRef]
  52. Fu, Y.; Viraraghavan, T. Removal of Congo Red from an Aqueous Solution by Fungus Aspergillus Niger. Adv. Environ. Res. 2002, 7, 239–247. [Google Scholar] [CrossRef]
  53. Wang, S.W.; Hines, A.L.; Farrier, D.S. Adsorption of Aliphatic Acids from Aqueous Solutions onto Activated Carbon. J. Chem. Eng. Data 1979, 24, 345–347. [Google Scholar] [CrossRef]
  54. Basha, S.; Murthy, Z.V.P.; Jha, B. Isotherm Modeling for Biosorption of Cu(II) and Ni(II) from Wastewater onto Brown Seaweed, Cystoseira Indica. AIChE J. 2008, 54, 3291–3302. [Google Scholar] [CrossRef]
  55. Chen, T.; Da, T.; Ma, Y. Reasonable Calculation of the Thermodynamic Parameters from Adsorption Equilibrium Constant. J. Mol. Liq. 2021, 322, 114980. [Google Scholar] [CrossRef]
  56. Tran, H.N.; Lima, E.C.; Juang, R.S.; Bollinger, J.C.; Chao, H.P. Thermodynamic Parameters of Liquid–Phase Adsorption Process Calculated from Different Equilibrium Constants Related to Adsorption Isotherms: A Comparison Study. J. Environ. Chem. Eng. 2021, 9, 106674. [Google Scholar] [CrossRef]
  57. Do, D.D. Adsorption Analysis: Equilibria and Kinetics; Series on Chemical Engineering; Imperial College Press: London, UK, 1998; Volume 2, ISBN 978-1-86094-130-6. [Google Scholar]
  58. Sayari, A.; Belmabkhout, Y.; Serna-Guerrero, R. Flue Gas Treatment via CO2 Adsorption. Chem. Eng. J. 2011, 171, 760–774. [Google Scholar] [CrossRef]
Figure 1. XRD pattern of the activated biocarbons and char derived from lemon peels under different treatment conditions.
Figure 1. XRD pattern of the activated biocarbons and char derived from lemon peels under different treatment conditions.
Molecules 29 04183 g001
Figure 2. N2 adsorption–desorption isotherms acquired at −196 °C for the activated biocarbons and char derived from lemon peels under different treatment conditions.
Figure 2. N2 adsorption–desorption isotherms acquired at −196 °C for the activated biocarbons and char derived from lemon peels under different treatment conditions.
Molecules 29 04183 g002
Figure 3. Pore size distribution of the activated biocarbons derived from lemon peels obtained on the basis of N2 (a) and CO2 (b) sorption.
Figure 3. Pore size distribution of the activated biocarbons derived from lemon peels obtained on the basis of N2 (a) and CO2 (b) sorption.
Molecules 29 04183 g003
Figure 4. SEM images of activated biocarbons and char derived from lemon peels.
Figure 4. SEM images of activated biocarbons and char derived from lemon peels.
Molecules 29 04183 g004
Figure 5. CO2 adsorption isotherms up to 1 bar at 0 °C for the activated biocarbons and char derived from lemon peels under different treatment conditions.
Figure 5. CO2 adsorption isotherms up to 1 bar at 0 °C for the activated biocarbons and char derived from lemon peels under different treatment conditions.
Molecules 29 04183 g005
Figure 6. Relationships between CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa, and specific surface area (SSA) (a), total pore volume (Vtot) (b), and micropore volume (Vmicro) (c).
Figure 6. Relationships between CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa, and specific surface area (SSA) (a), total pore volume (Vtot) (b), and micropore volume (Vmicro) (c).
Molecules 29 04183 g006
Figure 7. R2 values vs. narrow micropore diameter at a temperature of 0 °C and a pressure of 100 kPa.
Figure 7. R2 values vs. narrow micropore diameter at a temperature of 0 °C and a pressure of 100 kPa.
Molecules 29 04183 g007
Figure 8. CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa as a function of the volume of pores below 0.8594 (a), 0.8990 (b) and 0.8216 (c) nm.
Figure 8. CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa as a function of the volume of pores below 0.8594 (a), 0.8990 (b) and 0.8216 (c) nm.
Molecules 29 04183 g008
Figure 9. CO2 adsorption at temperatures of 0, 10, 20, and 30 °C onto activated biocarbon LP500_2 derived from lemon peels. The points represent experimental data; the lines were drawn using the fitted Radke–Prausnitz model.
Figure 9. CO2 adsorption at temperatures of 0, 10, 20, and 30 °C onto activated biocarbon LP500_2 derived from lemon peels. The points represent experimental data; the lines were drawn using the fitted Radke–Prausnitz model.
Molecules 29 04183 g009
Figure 10. Set of plots of ln(a) vs. T (a), b vs. 1/T (b), and n vs. 1/T (c), plotted for the calculation of ao, χ, Q, b0, n0, and α.
Figure 10. Set of plots of ln(a) vs. T (a), b vs. 1/T (b), and n vs. 1/T (c), plotted for the calculation of ao, χ, Q, b0, n0, and α.
Molecules 29 04183 g010
Figure 11. Adsorption isosteres with different surface coverage.
Figure 11. Adsorption isosteres with different surface coverage.
Molecules 29 04183 g011
Figure 12. Isosteric heat of CO2 adsorption on activated biocarbon LP500_2 derived from lemon peels calculated on the basis of adsorption measurements at temperatures of 0, 10, 20, and 30 °C.
Figure 12. Isosteric heat of CO2 adsorption on activated biocarbon LP500_2 derived from lemon peels calculated on the basis of adsorption measurements at temperatures of 0, 10, 20, and 30 °C.
Molecules 29 04183 g012
Table 1. Brief explanations of the activated biocarbons and char derived from lemon peels.
Table 1. Brief explanations of the activated biocarbons and char derived from lemon peels.
Sample NameThermal
Pretreatment
Ratio
KOH–Lemon
LP_1No pretreatment1
LP500_1 500 °C1
LP500_2500 °C2
LP500_4500 °C4
LP500500 °C-
Table 2. Textural parameters and CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa for activated biocarbons and char derived from lemon peels.
Table 2. Textural parameters and CO2 adsorption at a temperature of 0 °C and a pressure of 100 kPa for activated biocarbons and char derived from lemon peels.
ACSSA
(m2/g)
Vtot
(cm3/g)
Vmicro
(cm3/g)
Vmicro/Vtot
(%)
qCO2_0 °C
(mmol/g)
LP_110870.720.20282.01
LP500_1 18360.910.52584.99
LP500_221590.970.68705.69
LP500_428211.390.70504.97
LP5000.240.003--1.42
Table 3. Error functions values for the Freundlich, Langmuir, Toth, Sips, UNILAN, Radke–Prausnitz equations.
Table 3. Error functions values for the Freundlich, Langmuir, Toth, Sips, UNILAN, Radke–Prausnitz equations.
SSEHYBRIDAREMPSDSAE
CO2 adsorption at 0 °C
Freundlih0.0660.152.473.661.27
Langmuir0.220.414.135.592.43
Sips0.00230.00760.560.840.22
Toth0.000260.000670.170.240.17
UNILAN0.120.243.114.331.81
Radke–Prausnitz0.000180.000130.0580.0690.061
CO2 adsorption at 10 °C
Freundlih0.0380.102.183.300.89
Langmuir0.110.243.234.711.62
Sips0.0090.0150.681.000.54
Toth0.00820.0100.510.690.38
UNILAN0.0680.152.493.731.25
Radke–Prausnitz0.00800.00940.450.650.36
CO2 adsorption at 20 °C
Freundlih0.0290.0771.922.950.82
Langmuir0.0490.112.343.381.29
Sips0.0110.010.680.930.47
Toth0.0100.0120.570.750.44
UNILAN0.0330.0731.832.700.91
Radke–Prausnitz0.0100.0120.550.730.44
CO2 adsorption at 30 °C
Freundlih0.0190.0792.553.520.62
Langmuir0.0210.0732.443.180.69
Sips0.00570.0110.720.870.35
Toth0.00530.00820.500.620.30
UNILAN0.0150.0491.942.560.58
Radke–Prausnitz0.00520.00800.480.610.29
Table 4. The parameters of the Radke–Prausnitz model and errors of the model for activated biocarbon LP500_2 derived from lemon peels at different adsorption temperatures.
Table 4. The parameters of the Radke–Prausnitz model and errors of the model for activated biocarbon LP500_2 derived from lemon peels at different adsorption temperatures.
TempabnSSEHYBRIDAREMPSDSAE
0 °C0.1910.8020.5040.000180.000130.0580.0690.060
10 °C0.1400.6370.5180.00800.00940.460.650.36
20 °C0.0960.5410.5210.0100.0120.550.730.44
30 °C0.0680.4490.5300.00520.00800.470.610.29
Table 5. The parameters of temperature-dependent Radke–Prausnitz model equations.
Table 5. The parameters of temperature-dependent Radke–Prausnitz model equations.
ParameterValueUnit
Q13 090J/mol
b00.490kPa−1
n00.505
α0.24
a00.193mmol/g
χ10.49
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

Kiełbasa, K.; Siemak, J.; Sreńscek-Nazzal, J.; Benaouda, B.; Roy, B.; Michalkiewicz, B. Carbon Dioxide Adsorption over Activated Biocarbons Derived from Lemon Peel. Molecules 2024, 29, 4183. https://doi.org/10.3390/molecules29174183

AMA Style

Kiełbasa K, Siemak J, Sreńscek-Nazzal J, Benaouda B, Roy B, Michalkiewicz B. Carbon Dioxide Adsorption over Activated Biocarbons Derived from Lemon Peel. Molecules. 2024; 29(17):4183. https://doi.org/10.3390/molecules29174183

Chicago/Turabian Style

Kiełbasa, Karolina, Joanna Siemak, Joanna Sreńscek-Nazzal, Bestani Benaouda, Banasri Roy, and Beata Michalkiewicz. 2024. "Carbon Dioxide Adsorption over Activated Biocarbons Derived from Lemon Peel" Molecules 29, no. 17: 4183. https://doi.org/10.3390/molecules29174183

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop