Next Article in Journal
Influence of Mineral Additives on Strength Properties of Standard Mortar
Previous Article in Journal
Assessment of Microstructural Features of a Silchrome 1 Exhaust Valve of a Harley-Davidson WLA World War II Motorcycle
Previous Article in Special Issue
Effect of Thermal Activation on the Structure and Electrochemical Properties of Carbon Material Obtained from Walnut Shells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Activated Carbon for CO2 Adsorption from Avocado Seeds Activated with NaOH: The Significance of the Production Method

by
Joanna Siemak
1,
Grzegorz Mikołajczak
2,
Magdalena Pol-Szyszko
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, Piastów Ave. 42, 71-065 Szczecin, Poland
2
Faculty of Electrical Engineering, West Pomeranian University of Technology in Szczecin, 26 Kwietnia St. 10, 71-126 Szczecin, Poland
3
Department of Plant Genetics, Breeding and Biotechnology, Faculty of Environmental Management and Agriculture, West Pomeranian University of Technology in Szczecin, 17 Słowackiego Str., 71-434 Szczecin, Poland
*
Author to whom correspondence should be addressed.
Materials 2024, 17(16), 4157; https://doi.org/10.3390/ma17164157
Submission received: 14 July 2024 / Revised: 19 August 2024 / Accepted: 20 August 2024 / Published: 22 August 2024
(This article belongs to the Special Issue Progress in Carbon-Based Materials)

Abstract

:
The rise in atmospheric greenhouse gases like CO2 is a primary driver of global warming. Human actions are the primary factor behind the surge in CO2 levels, contributing to two-thirds of the greenhouse effect over the past decade. This study focuses on the chemical activation of avocado seeds with sodium hydroxide (NaOH). The influence of various preparation methods was studied under the same parameters: carbon precursor to NaOH mass ratio, carbonization temperature, and nitrogen flow. For two samples, preliminary thermal treatment was applied (500 °C). NaOH was used in the form of a saturated solution as well as dry NaOH. The same temperature of 850 °C of carbonization combined with chemical activation was applied for all samples. The applied modifications resulted in the following textural parameters: specific surface area from 696 to 1217 m2/g, total pore volume from 0.440 to 0.761 cm3/g, micropore volume from 0.159 to 0.418 cm3/g. The textural parameters were estimated based on nitrogen sorption at −196 °C. The XRD measurements and SEM pictures were also performed. CO2 adsorption was performed at temperatures of 0, 10, 20, and 30 °C and pressure up to 1 bar. In order to calculate the CO2 selectivity over N2 nitrogen adsorption at 20 °C was investigated. The highest CO2 adsorption (4.90 mmol/g) at 1 bar and 0 °C was achieved.

1. Introduction

The issue of global warming has been a worldwide priority for years, yet efforts to mitigate its effects remain insufficient. Recently, heightened attention to this issue has underscored that climate change, driven by greenhouse gas emissions, is raising alarm and making many people feel increasingly threatened [1,2]. Carbon dioxide is the most significant human-produced greenhouse gas, contributing to 77% of the anthropogenic greenhouse effect over the past decade, which translates to 26–30% of all CO2 emissions [3]. The primary source of human-related CO2 emissions is the burning of fossil fuels. The concentration of CO2 in flue gases varies depending on the type of fuel used; for example, coal results in 12–15 mol % CO2. In industries such as petroleum refining and other manufacturing processes, the CO2 levels in exhaust gases are process-dependent. For instance, oil refining emits 8–9 mol % CO2, cement production releases 14–33 mol % CO2, and iron and steel manufacturing results in 20–44 mol % CO2 [4].
There are multiple approaches to decreasing atmospheric CO2 emissions, such as the following: (a) improving the energy efficiency of fuels; (b) replacing fossil fuels with renewable energy sources like wind, solar, biomass, and geothermal power; (c) transitioning to technologies that use low-carbon energy sources such as natural gas; (d) reducing overall energy consumption; and (e) implementing pre-combustion, post-combustion, and oxy-combustion CO2 capture scenarios [5].
Additionally, more environmentally friendly and sustainable methods for CO2 capture have been explored, including the use of porous materials, especially carbon-based materials [6,7,8]. Carbon-based materials are renowned for their exceptional properties, such as adjustable and modifiable functionality, high surface area and total pore volume, especially micropore volume, broad temperature range for CO2 adsorption, rapid sorption kinetics, and excellent regeneration capabilities. In addition, these materials are cost-effective, industrially viable, and environmentally friendly [9,10]. They can be derived from a wide range of sources, and it is particularly worth highlighting that they can come from waste, especially from the food or agricultural industry [11,12].
Good CO2 sorbents prepared from a carbon source require chemical activation. KOH is mainly applied as the activation agent. Sometimes H3PO4, K2CO3, ZnCl2, and CaCl2 are also used to activate carbon production. NaOH is used very rarely in comparison to KOH. [11]. The reason is that usually activated carbons produced using NaOH showed much lower CO2 adsorption than activated carbons produced using KOH [13].
Nurfarhana et al. [14] used NaOH to treat natural rubber, converting it into activated carbon for CO2 adsorption. Natural rubber was carbonized at 400 °C for 2 h in a muffle furnace under a nitrogen flow. The pre-carbonized material was then mixed with solid NaOH at various ratios, ranging from 1:1 to 1:4 (natural rubber to NaOH). The activation process occurred in a tube furnace at 800 °C under a nitrogen atmosphere. Among all the samples, the highest specific surface area and total pore volume were achieved with a natural rubber to NaOH ratio of 1:4, reaching 1670 m2/g and 1.01 cm3/g, respectively. However, the CO2 adsorption for the mass ratio of 1:1 was relatively low, with a value of 0.69 mmol/g at 25 °C and 1 bar pressure. In contrast, the CO2 adsorption for the activated carbon obtained with a 1:4 ratio significantly increased to 2.98 mmol/g.
Dehkordi et al. [15] impregnated coal with NaOH solutions of varying concentrations (0.01–8 M) over different durations (4–7 h). Following impregnation, the coal samples were filtered and dried in an oven at 110 °C for 4–7 h. To examine the impact of washing on the activated carbon, some samples were immersed in deionized water. This study aimed to explore how different preparation parameters affect the surface properties and CO2 adsorption capacity of the activated carbon. The ideal conditions for good CO2 sorbent (51.4 mg/g) involved impregnation with a 1 M NaOH solution, followed by 6 h of mixing using a magnetic stirrer, without subsequent washing. The drying duration of the activated carbon had a minimal impact on its adsorption capacity.
Sodium hydroxide was used in the synthesis of sludge-activated carbon [16]. The carbon source was treated with a 5 wt % NaOH solution for 24 h. Following this, the reaction mixture was suction-filtered until the pH matched that of distilled water. Compared to the starting material, the specific surface area and pore volume of the sample treated with NaOH increased by 162% and 148%, respectively. The CO2 adsorption at a pressure of 4 MPa reached 65 cm3/g.
Solid monolithic graphene oxides (MGOs) were also used in the production of porous carbon using NaOH [17]. Graphene oxide (GO) was transformed into MGOs through a self-assembled reduction process at 90 °C, employing various weight ratios of oxalic acid (1:1, 1:0.500, and 1:0.250). Following this, the monoliths underwent carbonization at 600 °C and were chemically activated using different NaOH ratios (1:1, 1:2, and 1:3).
Increasing the mass ratios of MGO to NaOH from 1:1 to 1:2 led to a surface area increase of approximately 2.6 times, ranging from 520.8 to 753.9 m2/g (the surface area of untreated MGO was 289.2 m2/g). As a result, this significantly boosted the CO2 capture capacity to 2.10 mmol/g at a temperature of 25 °C and pressure of 1 bar.
Table 1 compares the activated carbon prepared using NaOH and standard materials (graphene and silica gel), focusing on the preparation method, properties, and CO2 adsorption capacities, with published literature.
Nitrogen adsorption studies at −196 °C and carbon dioxide adsorption in the temperature range of 0–30 °C form the basis of our research. Therefore, we present the theoretical foundations about isotherms and hysteresis loops as outlined in the fundamental literature on this subject, recommended by IUPAC [21,22].
Physisorption isotherms can generally be categorized into six types [21]. Type I isotherms are characteristic of microporous solids and are concave to the p/p0 axis, approaching a limiting value as p/p0 approaches 1. Type II isotherms are typical of non-porous or macroporous adsorbents and are reversible. Type III isotherms are less common, convex to the p/p0 axis throughout, and occur when adsorbate–adsorbate interactions are significant. Type IV isotherms are associated with mesoporous materials and feature a hysteresis loop due to capillary condensation within the mesopores and a limiting uptake at high p/p0. The initial section of a Type IV isotherm is indicative of monolayer–multilayer adsorption. Type V isotherms resemble Type III but include a hysteresis loop, observed in certain porous adsorbents with weak adsorbent–adsorbate interactions. Finally, Type VI isotherms represent stepwise multilayer adsorption on a uniform non-porous surface, with the sharpness of the steps varying depending on the system and temperature.
In the 1985 IUPAC recommendations, physisorption isotherms were categorized into six types. However, in 2015, Type I was further subdivided into Types Ia and Ib. Type Ia is characterized by a very steep uptake at very low p/p0, whereas for Type Ib, the uptake at very low p/p0 is much milder. Type Ia isotherms are typical of microporous materials with predominantly narrow micropores (diameter smaller than 1 nm). In contrast, Type Ib isotherms are observed in materials with a broader pore size distribution, including wider micropores and possibly narrow mesopores (diameter smaller than 2.5 nm).
In the original 1985 IUPAC classification, four types of hysteresis loops were identified [21]. This classification was expanded to six types in 2015 [22]. The shape of hysteresis loops is closely linked to specific characteristics of the pore structure and the underlying adsorption mechanism.
In Type H1, the two branches are almost vertical and nearly parallel over a significant range of gas uptake. This loop is typically found in materials with a narrow range of uniform mesopores. Additionally, Type H1 hysteresis can appear in networks of ink-bottle pores where the neck size distribution is similar to the pore/cavity size distribution. Type H2(a) loops are characterized by a very steep desorption branch. This feature can be attributed to pore-blocking/percolation in a narrow range of pore necks or to cavitation-induced evaporation. Type H2(b) loops are also associated with pore blocking, but the neck size distribution is much broader. H2(a) loops are commonly found in silica gels, some porous glasses, and certain ordered mesoporous materials. The Type H3 loop resembles a Type II isotherm on the adsorption branch, and the lower limit of the desorption branch is usually at the cavitation-induced p/p0. This loop occurs in non-rigid aggregates of plate-like particles and in pore networks with macropores that are not completely filled with pore condensate. Type H4 loops have nearly horizontal and parallel branches over a wide range of p/p0. This loop is associated with the filling of micropores and is found in micro–mesoporous carbons. Type H5 loops are unusual and have a distinctive form associated with pore structures containing both open and partially blocked mesopores.
When using chemical activation, activated carbons are typically produced without temperature pretreatment. However, if temperature pretreatment is applied, authors do not compare these two methods. Similarly, the activating agent is commonly used in solution form, although the dry form is rarely used. No one has compared materials produced using the activating agent in both dry and dissolved forms until now.
The international market for avocado fruit and its derivates (primarily guacamole and avocado oil) has been growing rapidly in the last decade [23]. Only the avocado pulp is employed for commercial applications, while other fruit elements like the seed and peel have no practical use and are disposed of by landfilling. Avocado seeds, which represent up to 26 wt % of the fruit mass, are produced in large amounts in centralized avocado transformation plants. Nevertheless, the use of this carbon source for producing activated carbon has been described in only a few publications. Therefore, we decided to use avocado seeds for producing carbon materials for CO2 adsorption. Our team is the only one investigating the potential of such utilization of these activated carbons.
Our studies concerned five activated carbons obtained using the same temperature (850 °C) of activation combined with carbonization and an equal mass ratio of carbon source to activator (1:1). NaOH was utilized in two different forms: solid NaOH and its saturated solution. Activated carbons were produced with and without preliminary carbonization. We used the relatively uncommon activator NaOH. KOH is commonly used. The reason is that usually activated carbons produced using NaOH showed much lower CO2 adsorption than activated carbons produced using KOH. We demonstrate here that CO2 adsorption using NaOH can also be high if the appropriate method is applied. Avocado seeds were used as the carbon source. Avocado seeds as a carbon source have been described in only a few publications, and only our team has successfully produced activated carbons from avocado seeds for effective CO2 adsorption. The novelty of the work lies in comparing the obtained materials using dry NaOH and a saturated solution as the activating agent, as well as highlighting the significance of preliminary carbonization. We demonstrated the significant importance of both the form of NaOH and the use of preliminary carbonization. These factors influence both the properties of activated carbons and their CO2 adsorption capacity.

2. Materials and Methods

2.1. Materials and Reagents

We used the following chemicals for the synthesis of activated carbon from avocado seeds: NaOH (Chempur, Piekary Śląskie, Poland, pure p.a.) and 35–38% hydrochloric acid (Chempur, pure p.a.). The avocado seeds were dried at 60 °C for 24 h before undergoing pretreatment at 500 °C and chemical activation. After drying, the seeds were crushed and powdered using a Royal RCMZ-800 (Royal Catering expondo GmbH, Berlin, Germany) multi grinder. The powdered avocado seeds were then used for the production of activated carbon.

2.2. Activated Carbon Synthesis

Five activated carbons were prepared from avocado seeds as described in Table 2 and shown in Figure 1.
In the case of the first two samples, an initial pretreatment of avocado seed was conducted at 500 °C. Subsequently, samples were mixed with a saturated solution of NaOH (C500_NaOH) or dry NaOH (C500_NaOHdry), followed by carbonization at a temperature of 850 °C for 1 h under nitrogen flow. They were then washed with 1 M HCl and rinsed with distilled water until reaching neutral pH. The two other samples were obtained as described above but without preliminary thermal treatment. The dried, powdered avocado seeds were mixed with a saturated solution of NaOH (C_NaOH) and dry NaOH (C_NaOHdry). The sample C_NaOHdry+H2O was obtained by mixing dried avocado seeds with dry NaOH and then adding some drops of water until the same pore volume of the starting materials was filled (incipient wetness method). In the incipient wetness method, only enough saturated NaOH solution is supplied to fill the pores of the starting materials. This minimizes the drying time of the sample, which reduces production costs. These costs are significantly lower than in the case of mixing the starting material with the saturated solution. In the latter method, the amount of saturated solution is such that the mass ratio of the carbon source to NaOH is 1. Here, we are dealing not only with the filling of the pores with the saturated solution.

2.3. Sample Characterization

The textural properties, including specific surface area (SSA), total pore volume (Vtot), micropore volume (Vmicro), and pore size distribution in the range of 0.3 to 30 nm, were assessed using N2 sorption at −196 °C. These measurements were conducted using an automatic volumetric adsorption device, the ASAP Sorption Surface Area and Pore Size Analyzer (ASAP 2460, Micrometrics, Novcross, GA, USA). The specific surface area was determined using the Brunauer–Emmett–Teller (BET) theory. The BET equation is useful for determining the surface area of nonporous materials as well as those with macropores and mesopores (having sufficiently large pores) within the relative pressure range of p/p0 from 0.05 to 0.3. However, this method is not suitable for microporous adsorbents. Because the materials were mostly microporous, the method suggested by Rouquerol et al. [24] was applied. The procedure relies on two criteria: (a) the constant C must be positive, and (b) the use of the BET equation must be confined to the range where the term V(1-p/p0) consistently increases with p/p0. The relative range for surface area calculation was selected individually for each sample, also considering the value of the correlation coefficient, and ranged between 5 × 10−5 and 1 × 10−2 for microporous materials and ranged between 1 × 10−2 and 2 × 10−1 C_NaOHdry.
The total pore volumes were determined from the volume of nitrogen adsorbed at a relative pressure close to p/p0 ≈ 1. The pore size distribution and micropore volumes were estimated using the density functional theory (N2-DFT model for slit pores). Equilibrium adsorption isotherms of carbon dioxide at temperatures ranging from 0 to 30 °C and nitrogen at 20 °C were measured volumetrically using the ASAP Sorption Surface Area and Pore Size Analyzer (ASAP 2460, Micrometrics, Novcross, GA, USA), with pressures up to 1 bar. We select CO2 adsorption at a temperature range of 0–30 °C because this range is most popular in the investigations. We can easily compare our results with those of other researchers.
Phase analysis of the activated carbons was conducted using an X-ray diffractometer equipped with a Cu lamp by PANalytical Empyrean.
The average graphene sheet diameter (La) and average crystallite thickness (Lc) were approximately calculated using the Scherrer equation [25]. Liu et al. [26] suggested K = 1.84 for the average graphene sheet diameter and K = 0.89 for the average crystallite thickness. The number of graphitic layers (N) was estimated using Equation (1):
N = L c d 002
The structural morphology of the activated carbons was examined using ultra-high-resolution field–emission scanning electron microscopy (UHR FE-SEM, Hitachi SU8020 Hitachi Ltd., Tokyo, Japan). The powdered sample was mounted on the SEM stub using double-sided carbon tape. Images were captured with a 20 kV accelerating voltage using a triple detector system.

3. Results and Discussion

3.1. Materials Characterization

The pore structure of activated carbons produced from avocado seeds was thoroughly investigated using N2 adsorption–desorption isotherms at −196 °C. As depicted in Figure 2, all adsorption isotherms exhibit a rapid N2 uptake at low pressure (relative pressure below 0.01), suggesting the presence of micropores in all the samples. The isotherms of samples pretreated at 500 °C and the sample prepared using NaOH solution show a sharp knee and a plateau after rapid uptake, indicating microporous materials. The isotherms of C500_NaOH and C500_NaOHdry correspond to a typical Type I isotherm according to the IUPAC classification [21], indicating that they are predominantly microporous. The very narrow hysteresis of H4 type is also present, indicating the presence of slit pores. For C_NaOH, we observed a mixed isotherm of Types I and II. This indicates the presence of micropores as well as macropores. For C_NaOH at higher relative pressure (>0.9), a significant increase in N2 adsorption is visible, which is evidence of the presence of macropores. The combined isotherms Types IV and II were identified for NaOHdry+H2O, C_NaOHdry. Type IV is characteristic of mesoporous materials. The very rapid uptake (Type II) at higher relative pressure (>0.9) indicates the presence of macropores. All the samples reveal hysteresis loops H4, which prove the presence of narrow slit-like pores. The hysteresis loop usually points to the presence of mesopores, but narrow hysteresis combined with a Type I isotherm indicates also microporosity.
The results of the pore size distribution calculations performed using the DFT method, presented in Figure 3, and the values of the textural parameters in Table 3 confirm the conclusions drawn from the shapes of the N2 adsorption isotherms. The pore size distribution curves (Figure 3) verify the microporous nature of C500_NaOHdry, C500_NaOH, and C_NaOH. The most microporosity resulted in significantly narrower and smaller ultramicropores (>0.7 nm). was observed for C500_NaOHdry. In contrast, C_NaOHdry and C_NaOHdry+H2O exhibited much wider and larger pores ranging from 2 to 4 nm.
Based on Table 3, it can be discerned that the obtained specific surface areas range from 696 to 1217 m2/g, total pore volume ranges from 0.440 to 0.761 cm3/g, micropore volume ranges from 0.159 to 0.418 cm3/g, and the ratio of the two mentioned parameters ranges from 22.08% to 76.42%. Despite using the same carbonization temperature and quantity of activating agent, the samples obtained differed according to the preparation methods used. The highest Vmicro/tot values are observed in the carbons that underwent preliminary thermal treatment of the seeds: C500_NaOHdry (76.42%) and C500_NaOH (65.91%).
It has been found that using a saturated NaOH solution as an activator for dry avocado seeds is more advantageous for developing porosity compared to using dry NaOH. The material activated with a saturated NaOH solution possessed a high specific surface area and micropore volume. The use of dry NaOH and dry avocado seeds was decidedly unfavorable. The situation was reversed when the activation was preceded by an initial pretreatment at a temperature of 500 °C. Activation of the resulting char with a saturated NaOH solution allowed for the production of a material with porosity comparable to that of C_NaOH obtained without pretreatment. The textural parameters of C500_NaOH were even lower. On the other hand, the activation of char after the pretreatment with dry NaOH enabled the production of a highly porous material with high micropore volume and specific surface area.
Based on theoretical considerations, the possible reaction between NaOH, carbon, and byproducts can be proposed as follows:
  • 4 NaOH + C → 4 Na + CO2 + 2 H2O
  • 6 NaOH + 2 C → 2 Na + 3 H2 + Na2CO3
  • 4 NaOH + 2 CO2 → 2 Na2CO3 + 2 H2O
  • Na2CO3 → Na2O + CO2
  • 4 Na2O + C → 4 Na + CO2
  • 4 NaOH + C → Na2CO3 + Na2O+ 2 H2
  • Na2CO3 +2 C→ 2 Na + 3 CO
  • C + CO2 ⇄ 2 CO
During the activation process, gases (H2, CO2, and CO), sodium, and sodium compounds (Na2O and Na2CO3). The reaction between the activating agent or byproducts and the carbon precursor leads to the breakdown of volatile organic compounds, resulting in the porous structure of the activated carbon samples. The gasification that occurs during activation with appropriate activating agents is crucial for the formation of pores. Thus, it can be argued that thermal treatment with an appropriate activating agent is essential for the formation of pores in activated carbon samples. During this process, the activating agent facilitates the development of a porous structure by promoting the removal of volatile substances and the creation of cavities. Without this step, the activated carbon would not exhibit the desired high surface area and porosity. Of course, both the choice of activating agent and the precise thermal conditions are crucial for producing high-quality activated carbon.
On the SEM images (Figure 4) of activated carbons with the highest specific surface area, such as C500_NaOHdry and C_NaOH, numerous macropores are observed, which, as commonly known, branch out into meso- and micropores within the material. For the material with the highest specific surface area (C500_NaOHdry), macropores have significantly smaller diameters on the surface than those in C_NaOH. In the case of other activated carbons with lower specific surface areas, the presence of thin, irregular planes creating occasional macropores with much larger sizes than in C_NaOH was observed. SEM images confirm the results obtained based on nitrogen adsorption measurements at a temperature of −196 °C.
Figure 5 illustrates the XRD pattern of activated carbons activated by NaOH. Two broad peaks appearing around 22 and 44 degrees are observable. The shapes of these peaks suggest that the activated carbons derived from avocado seeds were predominantly amorphous.
The average crystallite thickness, number of graphitic layers, and average graphene sheet diameter calculated on the basis of XRD measurements are listed in Table 4. The average crystallite thickness of samples pretreated at 500 °C is the lowest, so the average number of layers in a stack is about 2. The average crystallite thickness of samples produced using dry NaOH without pretreatment is the highest. The average number of layers in a stack for these samples and for C_NaOH are about 3. The average graphene sheet diameters of samples produced using dry NaOH without pretreatment are the highest, while C_NaOH is the lowest. The results obtained from XRD studies are consistent with the porosity analysis. The greatest disorder (the smallest average graphene sheet diameter) was observed for materials with the highest specific surface area.
CO2 adsorption studies were conducted at four different temperatures: 0, 10, 20, and 30 °C and at a pressure of 1 bar. The results are presented in Table 5 and Figure 6.
Every sample exhibits a reduction in CO2 adsorption with rising temperature, which is attributed to physisorption.
The CO2 isotherms, measured up to 1 bar at temperatures of 0–30 °C, show that C500_NaOHdry performs best at pressure below 1 bar. These observations can be rationalized by noting that C500_NaOHdry has the largest volumes of micropores, which are effective in adsorbing CO2 at pressure below 1 bar. The lowest CO2 adsorption was observed in activated carbons with the smallest micropore volume (C_NaOHdry and C_NaOHdry+H2O). The relationship between CO2 adsorption and micropore volume for all temperatures is presented in Figure 7. The micropore volume exhibited a very close correlation with CO2 adsorption, having R2 values higher than 0.9. This outcome aligns with findings from recent research studies [27,28].
The CO2 selectivity over N2 nitrogen adsorption at a temperature of 20 °C was calculated. The adsorption of nitrogen at a temperature of 20 °C was very low compared to the adsorption of CO2 at the same temperature (Figure 8).
Furthermore, the Ideal Adsorption Solution Theory (Equation (2)) was employed to calculate the theoretical selectivity of CO2 adsorption over N2 at a pressure of 1 bar and an equimolar composition. This approach is widely utilized to forecast the adsorption of individual components within gas mixtures based on data from single gas adsorption experiments.
S I A S T = q C O 2 @ p C O 2 / q N 2 @ p N 2 p C O 2 / p N 2
where qj @ pi is the adsorption capacity of i at pressure pi.
For equimolar composition, Equation (2) simplifies to Equation (3).
S C O 2 / N 2 = q C O 2 ( p ) q N 2 ( p )
The results of selectivity of CO2 adsorption over N2 for equimolar composition (Seq) at 1 bar are presented in Table 5. The selectivity values are quite high and vary from 6.76 to 8.05.
Based on CO2 and N2 adsorption measurements, it is possible to determine the selectivity of CO2 adsorption over N2 using the IAST method across the entire pressure range. To achieve this, it is necessary to establish an adsorption model that best fits the experimental data. The adsorption model also allows for the calculation of the isosteric heat of CO2 adsorption, which is extremely important.
The two-parameter (Langmuir [29] and Freundlich [30]) and three-parameter (Sips [31], Toth [32], and Radke–Prausnitz [33]) models were employed to fit the experimental data. To evaluate the appropriateness of fitting isotherm models to the experimental data, the Hybrid Error Function (HYBRID) [34] was utilized. The sum of squared errors (SSE) is commonly employed to measure the disparity between observed data and the true mean. To enhance SSE at lower pressure values, each square of the error values was divided by the experimental CO2 adsorption value.

3.2. Studies on C500_NaOHdry as Potential CO2 Adsorbent

Among all the models, the Radke–Prausnitz model provided the highest accuracy in estimations for CO2 adsorption, with the Hybrid Error Function not higher than 0.001. For N2 adsorption, the best fit was ascertained for the Toth model with HYBRID equal to 0.00011. The constants of the Radke–Prausnitz and Toth models for the best adsorbent (C500_NaOHdry) are collected in Table 6.
The Radke–Prausnitz equation is formulated as follows:
q = q m R P · b R P · p 1 + b R P · p n R P [ m m o l / g ]
  • qmRP—the maximum adsorption capacity [mmol/g];
  • bRP—the Radke–Prausnitz constant [bar−1];
  • nRP—Radke–Prausnitz model exponent.
Equation (5) outlines the Toth equation.
q = q m T b T p ( 1 + ( b T p ) n T ) 1 n T [ m m o l / g ]
  • qmT—the maximum adsorption capacity [mmol/g];
  • bT—the Toth constant [bar−1];
  • nT—the heterogeneity factor.
On the basis of the adsorption equilibrium models (2) and (3), the CO2 selectivity over N2 for equimolar composition at the pressure up to 1 bar for C500_NaOHdry was calculated and presented in Figure 9. The equimolar selectivity ranges from 19.17 to 6.84.
Post-combustion capture involves extracting CO2 from flue gas after combustion. Typically, the CO2 in flue gas is mixed with inert gases like nitrogen, argon, water vapor, and oxygen, resulting in a dilution of around 8–15% CO2. To determine if C500_NaOHdry is suitable as an adsorbent for CO2 removal from flue gas after combustion using the IAST method, the CO2 adsorption selectivity over N2 was calculated for CO2 contents of 8% and 15%. The selectivity for CO2 removal from flue gas is 16.65 and 13.48, respectively, which indicates that C500_NaOHdry is a potential sorbent for CO2 removal from flue gas.
The primary thermodynamic parameter of adsorption is the isosteric heat of adsorption. It can be determined from adsorption data at different temperatures by applying the Van’t Hoff or Clausius–Clapeyron Equation (6).
Q i s o = R l n ( p ) 1 T q
In order to apply Equation (6), the Radke–Prausnitz Equation (4) along with the parameters listed in Table 6 were used to compute the pressure values for the five specified surface coverage levels. Since it is not possible to solve Equation (4) algebraically for pressure, numerical methods were utilized instead. By plotting ln(p) against the inverse of the absolute temperature (1/T) for each loading, straight lines were produced, each having a slope of −Qiso/R. This method enables the calculation of the isosteric heat of adsorption for the five specified surface coverage levels (Figure 10).
A high isosteric heat of adsorption generally indicates stronger interactions between the adsorbate and the adsorbent. The isosteric heat of adsorption values are below 50 kJ/mol, suggesting that CO2 adsorption over C500_NaOHdry samples is primarily of a physical nature. The isosteric heat of adsorption up to 0.5 surface coverage remains essentially constant, ranging from 27.4 to 26.6 kJ/mol. Such consistent values are highly favorable for CO2 sorbent. CO2 molecules are bound to the C500_NaOHdry sorbent with moderate strength, allowing for desorption without excessive energy input. Materials with a higher isosteric heat of adsorption than C500_NaOHdry will require more energy to desorb CO2, making the desorption process more costly. It is important to emphasize that the heat of adsorption on C500_NaOHdry is extremely close to the heat of liquefaction of CO2 (25 kJ/mol), which is the lower limit for the isosteric heat of adsorption. Therefore, this material ensures the lowest adsorption costs.
Figure 11 shows adsorption–desorption isotherms at temperatures of 10 and 30 °C for C500_NaOHdry. The adsorption and desorption branches are nearly the same. The desorption isotherms lie slightly above the adsorption isotherms. Very narrow hysteresis is observed. The absorbed CO2 was completely released at low pressures, confirming the very weak interaction between CO2 and the sorbent [35].

4. Conclusions

This study focuses on the chemical activation of avocado seeds with NaOH. The influence of various preparation methods has been studied under the same parameters: carbon precursor to NaOH mass ratio, carbonization temperature, and nitrogen flow. We demonstrate here that CO2 adsorption using NaOH can be high if the appropriate method is applied.
The novelty of the work lies in comparing the obtained materials using dry NaOH and a saturated solution as the activating agent, as well as highlighting the significance of preliminary carbonization. We showed that preliminary carbonization at temperatures of 500 °C considerably improves CO2 uptake.
The significant importance of both the form of NaOH and the use of preliminary carbonization as parameters were demonstrated. The highest values of textural parameters and CO2 adsorption were achieved using thermal pretreatment and dry NaOH as the activating agent (C500_NaOHdry). Activation with NaOH enlarged the pores, resulting in an impressive specific surface area (1217 m2/g) and a significant micropore volume (0.547 cm3/g). This facilitated CO2 diffusion and adsorption (4.90 mmol CO2/g at 1 bar and 0 °C).
It was proven that the interaction between CO2 and C500_NaOHdry is very weak, and desorption is easy to achieve.
The equimolar CO2 selectivity over N2 was equal to 6.84, and for flue gas, it ranged from 16.65 to 13.48 depending on the CO2 content. The adsorbent exhibited quite low isosteric heat of adsorption, about 27 kJ/mol. All the values of selectivity and isosteric heat of adsorption indicate that C500_NaOHdry is a promising sorbent for CO2 capture from flue gas.
Of the various isotherm equations, the Radke–Prausnitz model proved effective in accurately predicting CO2 uptake isotherms.
In summary, the NaOH-activated carbon adsorbent exhibits potential for CO2 adsorption and separation applications, demonstrating impressive CO2 uptake and significant environmental benefits.

Author Contributions

Conceptualization, J.S. and B.M.; methodology, J.S. and B.M.; formal analysis, G.M.; investigation, J.S.; resources, M.P.-S.; writing—original draft preparation, J.S. and B.M.; writing—review and editing, J.S. and B.M.; visualization, J.S.; supervision, 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 raw data supporting the conclusions of this article will be made available by the authors upon request.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Li, S.; Yuan, X.; Deng, S.; Zhao, L.; Lee, K.B. A Review on Biomass-Derived CO2 Adsorption Capture: Adsorbent, Adsorber, Adsorption, and Advice. Renew. Sustain. Energy Rev. 2021, 152, 111708. [Google Scholar] [CrossRef]
  2. Song, C.; Liu, Q.; Ji, N.; Deng, S.; Zhao, J.; Li, Y.; Song, Y.; Li, H. Alternative Pathways for Efficient CO2 Capture by Hybrid Processes—A Review. Renew. Sustain. Energy Rev. 2018, 82, 215–231. [Google Scholar] [CrossRef]
  3. Vannak, H.; Osaka, Y.; Tsujiguchi, T.; Kodama, A. The Efficacy of Carbon Molecular Sieve and Solid Amine for CO2 Separation from a Simulated Wet Flue Gas by an Internally Heated/Cooled Temperature Swing Adsorption Process. Appl. Therm. Eng. 2024, 239, 122145. [Google Scholar] [CrossRef]
  4. Wei, R.; Alshahrani, T.; Chen, B.; Ibragimov, A.B.; Xu, H.; Gao, J. Advances in Porous Materials for Efficient Separation and Purification of Flue Gas. Sep. Purif. Technol. 2024, 352, 128238. [Google Scholar] [CrossRef]
  5. Rahimi, V.; Ferreiro-Salgado, A.; Gómez-Díaz, D.; Sonia Freire, M.; González-Álvarez, J. Evaluating the Performance of Carbon-Based Adsorbents Fabricated from Renewable Biomass Precursors for Post-Combustion CO2 Capture. Sep. Purif. Technol. 2024, 344, 127110. [Google Scholar] [CrossRef]
  6. Tang, D.; Lyu, X.; Huang, Z.; Xu, R.; Chen, J.; Qiu, T. Nitrogen-Doped Microporous Carbons as Highly Efficient Adsorbents for CO2 and Hg(II) Capture. Powder Technol. 2023, 427, 118769. [Google Scholar] [CrossRef]
  7. Dindi, A.; Quang, D.V.; Vega, L.F.; Nashef, E.; Abu-Zahra, M.R.M. Applications of Fly Ash for CO2 Capture, Utilization, and Storage. J. CO2 Util. 2019, 29, 82–102. [Google Scholar] [CrossRef]
  8. Xie, W.; Yao, X.; Li, H.; Li, H.; He, L. Biomass-Based N-Rich Porous Carbon Materials for CO2 Capture and In-situ Conversion. ChemSusChem 2022, 15, e202201004. [Google Scholar] [CrossRef]
  9. Sayari, A.; Belmabkhout, Y.; Serna-Guerrero, R. Flue Gas Treatment via CO2 Adsorption. Chem. Eng. J. 2011, 171, 760–774. [Google Scholar] [CrossRef]
  10. Sun, N.; Tang, Z.; Wei, W.; Snape, C.E.; Sun, Y. Solid Adsorbents for Low-Temperature CO2 Capture with Low-Energy Penalties Leading to More Effective Integrated Solutions for Power Generation and Industrial Processes. Front. Energy Res. 2015, 3, 9. [Google Scholar] [CrossRef]
  11. Ochedi, F.O.; Liu, Y.; Adewuyi, Y.G. State-of-the-Art Review on Capture of CO2 Using Adsorbents Prepared from Waste Materials. Process Saf. Environ. Prot. 2020, 139, 1–25. [Google Scholar] [CrossRef]
  12. Aghel, B.; Behaein, S.; Alobaid, F. CO2 Capture from Biogas by Biomass-Based Adsorbents: A Review. Fuel 2022, 328, 125276. [Google Scholar] [CrossRef]
  13. Kiełbasa, K.; Bayar, Ş.; Varol, E.A.; Sreńscek-Nazzal, J.; Bosacka, M.; Miądlicki, P.; Serafin, J.; Wróbel, R.J.; Michalkiewicz, B. Carbon Dioxide Adsorption over Activated Carbons Produced from Molasses Using H2SO4, H3PO4, HCl, NaOH, and KOH as Activating Agents. Molecules 2022, 27, 7467. [Google Scholar] [CrossRef]
  14. Nurfarhana, M.M.; Asikin-Mijan, N.; Yusoff, S.F.M. Porous Carbon from Natural Rubber for CO2 Adsorption. Mater. Chem. Phys. 2023, 308, 128196. [Google Scholar] [CrossRef]
  15. Dehkordi, S.S.R.; Delavar, Q.; Ebrahim, H.A.; Partash, S.S. CO2 Adsorption by Coal-Based Activated Carbon Modified with Sodium Hydroxide. Mater. Today Commun. 2022, 33, 104776. [Google Scholar] [CrossRef]
  16. Ma, J.; Liu, Y.; Chen, S.; Du, Y.; Wu, H. Changes in the Pore Structure of Modified Sludge-Activated Carbon and Its Effect on the Adsorption Characteristics of CO2 under High Pressure. Microporous Mesoporous Mater. 2022, 345, 112255. [Google Scholar] [CrossRef]
  17. Jha, R.K.; Bhunia, H.; Basu, S. Experimental Kinetics and Thermodynamics Investigation: Chemically Activated Carbon-Enriched Monolithic Reduced Graphene Oxide for Efficient CO2 Capture. Heliyon 2024, 10, e27439. [Google Scholar] [CrossRef]
  18. Tahmasebpoor, M.; Iranvandi, M.; Heidari, M.; Azimi, B.; Pevida, C. Development of Novel Waste Tea-Derived Activated Carbon Promoted with SiO2 Nanoparticles as Highly Robust and Easily Fluidizable Sorbent for Low-Temperature CO2 Capture. J. Environ. Chem. Eng. 2023, 11, 110437. [Google Scholar] [CrossRef]
  19. Singh, S.B.; De, M. Carbon Dioxide Removal by Chemically and Thermally Reduced Graphene-Based Adsorbents. Korean J. Chem. Eng. 2024, 41, 783–796. [Google Scholar] [CrossRef]
  20. Avellaneda, G.L.; Denoyel, R.; Beurroies, I. CO2/H2O Adsorption and Co-Adsorption on Functionalized and Modified Mesoporous Silicas. Microporous Mesoporous Mater. 2024, 363, 112801. [Google Scholar] [CrossRef]
  21. Sing, K.S.W. Reporting Physisorption Data for Gas/Solid Systems with Special Reference to the Determination of Surface Area and Porosity (Recommendations 1984). Pure Appl. Chem. 1985, 57, 603–619. [Google Scholar] [CrossRef]
  22. 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]
  23. Domínguez, M.P.; Araus, K.; Bonert, P.; Sánchez, F.; San Miguel, G.; Toledo, M. The Avocado and Its Waste: An Approach of Fuel Potential/Application. In Environment, Energy and Climate Change II; Springer: Cham, Switzerland, 2014; pp. 199–223. [Google Scholar]
  24. Rouquerol, J.; Llewellyn, P.; Rouquerol, F. Is the Bet Equation Applicable to Microporous Adsorbents? Stud. Surf. Sci. Catal. 2007, 160, 49–56. [Google Scholar]
  25. Cullity, B.D.; Weymouth, J.W. Elements of X-ray Diffraction. Am. J. Phys. 1957, 25, 394–395. [Google Scholar] [CrossRef]
  26. Lu, L.; Sahajwalla, V.; Kong, C.; Harris, D. Quantitative X-ray Diffraction Analysis and Its Application to Various Coals. Carbon 2001, 39, 1821–1833. [Google Scholar] [CrossRef]
  27. Singh, G.; Tiburcius, S.; Ruban, S.M.; Shanbhag, D.; Sathish, C.I.; Ramadass, K.; Vinu, A. Pure and Strontium Carbonate Nanoparticles Functionalized Microporous Carbons with High Specific Surface Areas Derived from Chitosan for CO2 Adsorption. Emergent Mater. 2019, 2, 337–349. [Google Scholar] [CrossRef]
  28. Serafin, J.; Sreńscek-Nazzal, J.; Kamińska, A.; Paszkiewicz, O.; Michalkiewicz, B. Management of Surgical Mask Waste to Activated Carbons for CO2 Capture. J. CO2 Util. 2022, 59, 101970. [Google Scholar] [CrossRef]
  29. Elmorsi, T.M. Equilibrium Isotherms and Kinetic Studies of Removal of Methylene Blue Dye by Adsorption onto Miswak Leaves as a Natural Adsorbent. J. Environ. Prot. 2011, 2, 817–827. [Google Scholar] [CrossRef]
  30. Travis, C.C.; Etnier, E.L. A Survey of Sorption Relationships for Reactive Solutes in Soil. J. Environ. Qual. 1981, 10, 8–17. [Google Scholar] [CrossRef]
  31. Ayawei, N.; Angaye, S.S.; Wankasi, D.; Dikio, E.D. Synthesis, Characterization and Application of Mg/Al Layered Double Hydroxide for the Degradation of Congo Red in Aqueous Solution. Open J. Phys. Chem. 2015, 5, 56–70. [Google Scholar] [CrossRef]
  32. Podder, M.S.; Majumder, C.B. Studies on the Removal of As(III) and As(V) through Their Adsorption onto Granular Activated Carbon/MnFe2O4 Composite: Isotherm Studies and Error Analysis. Compos. Interfaces 2016, 23, 327–372. [Google Scholar] [CrossRef]
  33. Tran, H.N.; Bollinger, J.-C.; Lima, E.C.; Juang, R.-S. How to Avoid Mistakes in Treating Adsorption Isotherm Data (Liquid and Solid Phases): Some Comments about Correctly Using Radke-Prausnitz Nonlinear Model and Langmuir Equilibrium Constant. J. Environ. Manag. 2023, 325, 116475. [Google Scholar] [CrossRef]
  34. Kumar, K.V.; Porkodi, K.; Rocha, F. Comparison of Various Error Functions in Predicting the Optimum Isotherm by Linear and Non-Linear Regression Analysis for the Sorption of Basic Red 9 by Activated Carbon. J. Hazard. Mater. 2008, 150, 158–165. [Google Scholar] [CrossRef] [PubMed]
  35. Sapchenko, S.A.; Barsukova, M.O.; Belosludov, R.V.; Kovalenko, K.A.; Samsonenko, D.G.; Poryvaev, A.S.; Sheveleva, A.M.; Fedin, M.V.; Bogomyakov, A.S.; Dybtsev, D.N.; et al. Understanding Hysteresis in Carbon Dioxide Sorption in Porous Metal–Organic Frameworks. Inorg. Chem. 2019, 58, 6811–6820. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Activated carbon preparation from avocado seeds.
Figure 1. Activated carbon preparation from avocado seeds.
Materials 17 04157 g001
Figure 2. N2 sorption isotherms measured at −196 °C, adsorption brunch—filled symbols, desorption brunch—empty symbols.
Figure 2. N2 sorption isotherms measured at −196 °C, adsorption brunch—filled symbols, desorption brunch—empty symbols.
Materials 17 04157 g002
Figure 3. Pore size distribution of activated carbon.
Figure 3. Pore size distribution of activated carbon.
Materials 17 04157 g003
Figure 4. SEM images (30,000× magnification).
Figure 4. SEM images (30,000× magnification).
Materials 17 04157 g004
Figure 5. XRD pattern of activated carbons.
Figure 5. XRD pattern of activated carbons.
Materials 17 04157 g005
Figure 6. CO2 adsorption isotherms at different temperatures.
Figure 6. CO2 adsorption isotherms at different temperatures.
Materials 17 04157 g006
Figure 7. Correlation between CO2 adsorption and micropore volume.
Figure 7. Correlation between CO2 adsorption and micropore volume.
Materials 17 04157 g007
Figure 8. N2 adsorption isotherms.
Figure 8. N2 adsorption isotherms.
Materials 17 04157 g008
Figure 9. The CO2 selectivity over N2 for equimolar composition for the pressure up to 1 bar for C500_NaOHdry.
Figure 9. The CO2 selectivity over N2 for equimolar composition for the pressure up to 1 bar for C500_NaOHdry.
Materials 17 04157 g009
Figure 10. Isosteric heat of adsorption over C500_NaOHdry versus surface coverage.
Figure 10. Isosteric heat of adsorption over C500_NaOHdry versus surface coverage.
Materials 17 04157 g010
Figure 11. CO2 sorption isotherms measured at 10 and 30 °C for C500_NaOHdry.
Figure 11. CO2 sorption isotherms measured at 10 and 30 °C for C500_NaOHdry.
Materials 17 04157 g011
Table 1. The activated carbons prepared using NaOH.
Table 1. The activated carbons prepared using NaOH.
SorbentMethod of PreparationSSA (m2/g)
Vtot (cm3/g)
qCO2_
(mmol/g)
t (°C)
p (bar)
Ref.
AC from natural rubbercarbonized at 400 °C; mixed with solid NaOH at ratios ranging from 1:1 to 1:4 (natural rubber to NaOH); activation at 800 °C1670 m2/g
1.01 cm3/g
2.9825
1
[14]
AC from coalimpregnation with NaOH solutions concentrations (0.01–8 M), durations (4–7 h).614 m2/g1.6825
2.12
[15]
AC from sludgetreating with a 5 wt % NaOH solution for 24 h; washing with water until the pH was neutral.101 m2/g
0.123 cm3/g
2.9030
40
[16]
AC from monolithic graphene oxidesmixed with NaOH at ratios ranging from 1 to 3; carbonized at 600 °C; washing with water until the pH was neutral.754 m2/g
1.97 cm3/g
2.1025
1
[17]
AC from waste teaimpregnation with solution of NaOH (mas ratio 1:1); carbonized at 600 °C; washing with water until the pH was neutral.270 m2/g
0.106 cm3/g
0.9625
1
[18]
Graphene-hydrazinegraphite oxide was treated by hydrazine hydrate at a temperature of 90 °C409 m2/g
0.48 cm3/g
1.440
1
[19]
SBA-15 midified by BTESETEOS was replaced by 10% of 3-(triethoxysilyl) propylamine.269 m2/g
0.36 cm3/g
1.2130
1
[20]
Table 2. Procedures of sample preparation.
Table 2. Procedures of sample preparation.
C500_NaOHavocado seeds powder → furnace 500 °C → NaOH solution → furnace 850 °C
C500_NaOHdryavocado seeds powder → furnace 500 °C → NaOH dry → furnace 850 °C
C_NaOHavocado seeds powder +NaOH solution → furnace 850 °C
C_NaOHdryavocado seeds powder +NaOH dry → furnace 850 °C
C_NaOHdry+H2Oavocado seeds powder +NaOH dry + few drops of H2O → furnace 850 °C
Table 3. Textural parameters of activated carbon.
Table 3. Textural parameters of activated carbon.
ACSSAVtotVmicroVmicro/tot
(m2/g)(cm3/g)(cm3/g)(%)
C_NaOH9180.5740.29551.39
C_NaOHdry7000.7160.15922.21
C_NaOHdry+H2O6960.7610.16822.08
C500_NaOH8850.4400.29065.91
C500_NaOHdry12170.5470.41876.42
Table 4. The average crystallite thickness (Lc), number of graphitic layers (N), and average graphene sheet diameter (La).
Table 4. The average crystallite thickness (Lc), number of graphitic layers (N), and average graphene sheet diameter (La).
ACLcNLa
(nm) (nm)
C_NaOH1.042.712.39
C_NaOHdry1.233.274.10
C_NaOHdry+H2O1.233.303.69
C500_NaOH0.882.293.45
C500_NaOHdry1.042.712.39
Table 5. CO2 and N2 adsorption values over activated carbon.
Table 5. CO2 and N2 adsorption values over activated carbon.
ACqCO2_0 °CqCO2_10 °CqCO2_20 °CqCO2_30 °CqN2_20 °CSeq
(mmol/g)(mmol/g)(mmol/g)(mmol/g)(mmol/g)
C_NaOH3.693.072.572.120.386.76
C_NaOHdry2.372.101.691.190.218.05
C_NaOHdry+H2O2.562.161.641.210.217.81
C500_NaOH4.203.502.962.460.378.00
C500_NaOHdry4.904.003.372.760.506.84
Table 6. The constants of Radke–Prausnitz and Toth models for C500_NaOHdry.
Table 6. The constants of Radke–Prausnitz and Toth models for C500_NaOHdry.
Radke–Prausnitz Model for CO2Toth Model for N2
Temperature [°C]
0102030 20
qmRP5.765.014.393.83qmT3.01
bRP5.814.393.332.61bT0.21
nRP0.590.590.590.6nT0.90
HYBRID 0.001030.000780.000280.00081HYBRID 0.00011
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

Siemak, J.; Mikołajczak, G.; Pol-Szyszko, M.; Michalkiewicz, B. Activated Carbon for CO2 Adsorption from Avocado Seeds Activated with NaOH: The Significance of the Production Method. Materials 2024, 17, 4157. https://doi.org/10.3390/ma17164157

AMA Style

Siemak J, Mikołajczak G, Pol-Szyszko M, Michalkiewicz B. Activated Carbon for CO2 Adsorption from Avocado Seeds Activated with NaOH: The Significance of the Production Method. Materials. 2024; 17(16):4157. https://doi.org/10.3390/ma17164157

Chicago/Turabian Style

Siemak, Joanna, Grzegorz Mikołajczak, Magdalena Pol-Szyszko, and Beata Michalkiewicz. 2024. "Activated Carbon for CO2 Adsorption from Avocado Seeds Activated with NaOH: The Significance of the Production Method" Materials 17, no. 16: 4157. https://doi.org/10.3390/ma17164157

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