Acessibilidade / Reportar erro

Ammonia Removal in Activated Carbons Prepared from Olive Oil Industry Waste

Abstract

Activated carbons (ACs) from olive stone were prepared using CO2, steam, KOH, and H3PO4 as activating agents. The resultant activated carbons were characterized by proximate and ultimate analysis, N2 adsorption (Brunauer-Emmett-Teller (BET) method), iodine number, Boehm titration, temperature-programmed desorption (TPD), and Fourier transform infrared spectroscopy (FTIR). Ammonia (NH3) was used as a test molecule to be adsorbed. The BET surface areas of the ACs obtained ranged from 1000 to 1986 m2 g-1. Type I isotherms were obtained for all the samples, although steam and H3PO4 ACs showed a significant mesopore contribution. KOH activation resulted in carbon with a high microporosity (98%) and high iodine adsorption (1030 mg g-1). KOH AC prepared with a KOH/pyrolyzed char weight ratio of 2 and at 900 °C showed the highest NH3 adsorption (252 mg g-1), favored by the high microporosity and adequate acidity. Chemical activation (KOH and H3PO4) promotes higher NH3 adsorption than the physical ACs prepared (CO2 and steam). Langmuir and Freundlich adsorption equilibrium models were used to correlate the NH3 adsorption isotherms, obtaining the best fit for the Freundlich equation. The results indicated that olive stone-based activated carbon could be used for commercial AC to remove NH3 from gaseous streams.

Keywords:
olive stones; activated carbon; physical and chemical activation; adsorption isotherms; ammonia removal


Introduction

Activated carbon (AC) is a well-known material employed to separate and remove unwanted substances present in industrial effluents. Because of its affordable price, high porosity, internal surface area, and excellent adsorptive capabilities, AC is the most extensively used adsorbent.11 Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref] This material is commonly used in the chemical and food industries, as well as other process industries (drinking water treatment, pharmaceuticals, etc.).22 Demiral, H.; Demiral, I.; Karabacakoĝlu, B.; Tümsek, F.; Chem. Eng. Res. Des. 2011, 89, 206. [Crossref]

Physical and chemical activations are employed to prepare activated carbons (ACs). In the physical method, the raw material is carbonized between 400 and 800 °C, followed by gasification at higher temperatures in an oxidizing gas flow (steam, CO2, air, or a combination).33 Chang, C.-F.; Chang, C.-Y.; Tsai, W.-T.; J. Colloid Interface Sci. 2000, 232, 45. [Crossref],44 Rodríguez-Reinoso, F.; Molina-Sabio, M.; Carbon N. Y. 1992, 30, 1111. [Crossref] Hence, the physical activation is a two-step process but can also be done in a one-step process.55 Yang, K.; Peng, J.; Xia, H.; Zhang, L.; Srinivasakannan, C.; Guo, S.; J. Taiwan Inst. Chem. Eng. 2010, 41, 367. [Crossref] Carbonization and activation are carried out simultaneously in chemical activation, using a chemical substance (e.g., H2SO4, H3PO4, ZnCl2, KOH, NaOH) and then subjected to a heat treatment (500-800 °C) in a one-step process.66 Demiral, H.; Demiral, I.; Fatma, T.; Karabacako, B.; Surf. Interface Anal. 2008, 40, 616. [Crossref],77 Ubago-Pérez, R.; Carrasco-Marín, F.; Fairén-Jiménez, D.; Moreno-Castilla, C.; Microporous Mesoporous Mater. 2006, 92, 64. [Crossref] The surface area, porosity, and adsorption properties of AC are determined by the type of the raw material, as well as the method and process parameters used for carbonization and activation.88 Román, S.; González, J. F.; González-García, C. M.; Zamora, F.; Fuel Process. Technol. 2008, 89, 715. [Crossref],99 González, J. F.; Román, S.; Encinar, J. M.; Martínez, G.; J. Anal. Appl. Pyrolysis 2009, 85, 134. [Crossref]

Anthracite, peat, lignite, bituminous coal, and lignocellulosic materials such as coconut shells and wood are among the most popular carbon-containing feedstocks employed as a precursor to producing AC.1010 Ahmedna, M.; Marshall, W. E.; Rao, R. M.; Bioresour. Technol. 2000, 71, 113. [Crossref] In the last decades, an increasing number of investigations related to the production of low-cost AC from agricultural wastes such as sugarcane bagasse,1010 Ahmedna, M.; Marshall, W. E.; Rao, R. M.; Bioresour. Technol. 2000, 71, 113. [Crossref] pistachio-nut shells,1111 Lua, A. C.; Yang, T.; Guo, J.; J. Anal. Appl. Pyrolysis 2004, 72, 279. [Crossref] corn cob,33 Chang, C.-F.; Chang, C.-Y.; Tsai, W.-T.; J. Colloid Interface Sci. 2000, 232, 45. [Crossref] date stones,1212 Bouchelta, C.; Medjram, M. S.; Bertrand, O.; Bellat, J.-P.; J. Anal. Appl. Pyrolysis 2008, 82, 70. [Crossref] neem husk,1313 Alau, K. K.; Gimba, C. E.; Kagbu, J. A.; Arch. Appl. Sci. Res. 2010, 2, 456. [Crossref] almond shell,99 González, J. F.; Román, S.; Encinar, J. M.; Martínez, G.; J. Anal. Appl. Pyrolysis 2009, 85, 134. [Crossref] fir bark,1414 Luo, L.; Wu, X.; Li, Z.; Zhou, Y.; Chen, T.; Fan, M.; Zhao, W.; R. Soc. Open Sci. 2019, 6, 190523. [Crossref] jackfruit,1515 Nagalakshmi, T. V.; Emmanuel, K. A.; Suresh Babu, C.; Chakrapani, C.; Divakar, P. P.; Int. Lett. Chem., Phys. Astron. 2015, 54, 189. [Crossref] cherry stones,1616 Olivares-Marín, M.; Fernández-González, C.; Macías-García, A.; Gómez-Serrano, V.; Appl. Surf. Sci. 2006, 252, 5980. [Crossref] coffee grounds,1717 Laksaci, H.; Khelifi, A.; Trari, M.; Addoun, A.; J. Cleaner Prod. 2017, 147, 254. [Crossref] rice husk,1818 Hanum, F.; Bani, O.; Izdiharo, A. M.; IOP Conf. Ser.: Mater. Sci. Eng. 2017, 180, 012149. [Crossref] peach stones,1919 Álvarez-Torrellas, S.; Peres, J. A.; Gil-Álvarez, V.; Ovejero, G.; García, J.; Chem. Eng. J. 2017, 320, 319. [Crossref] palm kernel shell,2020 To, M.-H.; Hadi, P.; Hui, C. W.; Lin, C. S. K.; McKay, G.; J. Mol. Liq. 2017, 241, 386. [Crossref] and olive stones,2121 Berrios, M.; Martín, M.; Martín, A.; J. Ind. Eng. Chem. 2012, 18, 780. [Crossref]

22 Yakout, S. M.; Sharaf El-Deen, G.; Arabian J. Chem. 2016, 9, S1155. [Crossref]
-2323 Yahya, M. A.; Al-Qodah, Z.; Ngah, C. W. Z.; Renewable Sustainable Energy Rev. 2015, 46, 218. [Crossref] have been reported. These ACs have displayed good performance and could replace expensive coal-based carbons in various industrial applications to remove different pollutants (e.g., dyes, metal ions, organic pollutants, pharmaceuticals and personal care products). The AC prepared from an abundant waste source would include economic gains and environmental benefits.

In this sense, the olive agribusiness in Chile has about 25,000 hectares of olive trees, with a production of 18,500 tons of oil in 2019 destined for the national and international market, ranking Chile in tenth place among the olive oil-exporting countries in the world.2424 Chile Oliva; Informe Anual Mercado Nacional de Aceite de Oliva; 2019. [Link] accessed in August 2022 The oil extraction process generates about 70,000 tons of solid waste, an organic residue with high humidity that is used directly on soils as fertilizer, compost, or fuel in the same process.2525 Foods from Chile; La Industria del Aceite de Oliva en Chile. Informe de Sustentabilidad Chile Oliva; Chile, 2019. [Link] accessed in August 2022

26 Fundación para la Innovación Agraria; Con Residuos del Aceite de Oliva Desarrollan Biocombustibles Sólidos y Carbones Descontaminantes; Chile, 2020. [Link] accessed in August 2022
-2727 Zabaniotou, A.; Stavropoulos, G.; Skoulou, V.; Bioresour. Technol. 2008, 99, 320. [Crossref] However, a large part of this waste is not used and is accumulated, generating environmental problems associated with foul odors.2828 Azbar, N.; Bayram, A.; Filibeli, A.; Muezzinoglu, A.; Sengul, F.; Ozer, A.; Crit. Rev. Environ. Sci. Technol. 2004, 34, 209. [Crossref]

The pyrolysis process of agricultural by-products has been considered a technology to solve the disposal problem associated with this waste.99 González, J. F.; Román, S.; Encinar, J. M.; Martínez, G.; J. Anal. Appl. Pyrolysis 2009, 85, 134. [Crossref],2323 Yahya, M. A.; Al-Qodah, Z.; Ngah, C. W. Z.; Renewable Sustainable Energy Rev. 2015, 46, 218. [Crossref],2929 Martínez, M. L.; Torres, M. M.; Guzmán, C. A.; Maestri, D. M.; Ind. Crops Prod. 2006, 23, 23. [Crossref]

30 Ioannidou, O.; Zabaniotou, A.; Renewable Sustainable Energy Rev. 2007, 11, 1966. [Crossref]
-3131 Wong, S.; Ngadi, N.; Inuwa, I. M.; Hassan, O.; J. Cleaner Prod. 2018, 175, 361. [Crossref] Using this process, it is possible to generate add-value products such as liquids and gases with fuel characteristics, and the char can be used as a precursor of the AC.99 González, J. F.; Román, S.; Encinar, J. M.; Martínez, G.; J. Anal. Appl. Pyrolysis 2009, 85, 134. [Crossref] The proposed technological solution aims to strengthen the olive industry in Chile through the integral recovery of its residues through the development of ACs, taking advantage of the intrinsic characteristics of the olive stone.

For this reason, the goal of this study is to evaluate the use of the olive stone to obtain ACs by physical (CO2 and steam) and chemical activation (KOH and H3PO4), testing them for NH3 adsorption and establishing which method can promote the best AC for the proposed application. Although olive stone as a raw material for obtaining AC has been widely studied,11 Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref],44 Rodríguez-Reinoso, F.; Molina-Sabio, M.; Carbon N. Y. 1992, 30, 1111. [Crossref],88 Román, S.; González, J. F.; González-García, C. M.; Zamora, F.; Fuel Process. Technol. 2008, 89, 715. [Crossref],2222 Yakout, S. M.; Sharaf El-Deen, G.; Arabian J. Chem. 2016, 9, S1155. [Crossref],3232 González, M. T.; Rodríguez-Reinoso, F.; García, A. N.; Marcilla, A.; Carbon N. Y. 1997, 35, 159. [Crossref]

33 Enaime, G.; Ennaciri, K.; Ounas, A.; Baçaoui, A.; Seffen, M.; Selmi, T.; Yaacoubi, A.; J. Mater. Environ. Sci. 2017, 8, 4125. [Crossref]

34 Moreno-Castilla, C.; Carrasco-Marín, F.; López-Ramón, M. V.; Alvarez-Merino, M. A.; Carbon N. Y. 2001, 39, 1415. [Crossref]

35 Rodríguez-Reinoso, F.; Molina-Sabio, M.; González, M. T.; Carbon N. Y. 1995, 33, 15. [Crossref]

36 Baçaoui, A.; Yaacoubi, A.; Dahbi, A.; Bennouna, C.; Phan Tan Luu, R.; Maldonado-Hodar, F. J.; Rivera-Utrilla, J.; Moreno-Castilla, C.; Carbon N. Y. 2001, 39, 425. [Crossref]

37 Linares-Solano, A.; Rodriguez-Reinoso, F.; Molina-Sabio, M.; Lopez-Gonzalez, J. D.; Adsorpt. Sci. Technol. 1984, 1, 223. [Crossref]

38 Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A. F.; Maldonado-Hódar, F. J.; Carrasco-Marín, F.; Electrochim. Acta 2017, 229, 219. [Crossref]
-3939 Molina-Sabio, M.; Rodríguez-Reinoso, F.; Colloids Surf., A 2004, 241, 15. [Crossref] few investigations have focused on simultaneously evaluating the four activations carried out in this work. Concerning NH3, this gas is one of the most commonly released compounds in the chemical industry, and its presence can be smelt in the air even at low concentrations (< 50 ppm), causing eye, throat, and nose irritation.4040 Mochizuki, T.; Kubota, M.; Matsuda, H.; D’Elia Camacho, L. F.; Fuel Process. Technol. 2016, 144, 164. [Crossref],4141 Petit, C.; Karwacki, C.; Peterson, G.; Bandosz, T. J.; J. Phys. Chem. C 2007, 111, 12705. [Crossref] The use of AC is a promising approach to removing this gas, considering its simplicity and economic operation.4242 Busca, G.; Pistarino, C.; J. Loss Prev. Process Ind. 2003, 16, 157. [Crossref] Besides, it has been reported that the acidic surface functionalities of AC favor NH3 adsorption and the direct interaction between the gas molecules and the carbon surface.4343 Guo, J.; Xu, W. S.; Chen, Y. L.; Lua, A. C.; J. Colloid Interface Sci. 2005, 281, 285. [Crossref] Therefore, the production of AC from the waste of olive oil production can become one of the solutions for the effective elimination of residues and contribute to the sustainability and competitiveness of small and medium-sized industries in this sector.

Experimental

Materials

Olive stones were supplied by a local olive oil manufacturer (Terramater, Curicó Valley-Region of Maule, Chile). These materials were processed without previous treatment, crushed and sieved until particle sizes between 0.6 and 1 mm were obtained. For comparison purposes, the C Gran and GAC 1240 Norit (Cabot Corporation, Massachusetts, USA) commercial ACs were included in the analysis (see Supplementary Information section).

Preparation of ACs

ACs were prepared using physical and chemical methods. Physical activation was done employing CO2 (Linde, 99.5%, Santiago, Chile) and steam, and for chemical activation, KOH (Merck, 85% purity, Darmstadt, Germany) and H3PO4 (Merck, 85% purity, Darmstadt, Germany) reagents were used. After reviewing some references, the conditions to pyrolyze and activate the material were considered.55 Yang, K.; Peng, J.; Xia, H.; Zhang, L.; Srinivasakannan, C.; Guo, S.; J. Taiwan Inst. Chem. Eng. 2010, 41, 367. [Crossref],77 Ubago-Pérez, R.; Carrasco-Marín, F.; Fairén-Jiménez, D.; Moreno-Castilla, C.; Microporous Mesoporous Mater. 2006, 92, 64. [Crossref],1010 Ahmedna, M.; Marshall, W. E.; Rao, R. M.; Bioresour. Technol. 2000, 71, 113. [Crossref],1111 Lua, A. C.; Yang, T.; Guo, J.; J. Anal. Appl. Pyrolysis 2004, 72, 279. [Crossref],2222 Yakout, S. M.; Sharaf El-Deen, G.; Arabian J. Chem. 2016, 9, S1155. [Crossref],3535 Rodríguez-Reinoso, F.; Molina-Sabio, M.; González, M. T.; Carbon N. Y. 1995, 33, 15. [Crossref],4444 Haimour, N. M.; Emeish, S.; Waste Manage. 2006, 26, 651. [Crossref]

45 Solum, M. S.; Pugmire, R. J.; Jagtoyen, M.; Derbyshire, F.; Carbon N. Y. 1995, 33, 1247. [Crossref]

46 Jagtoyen, M.; Derbyshire, F.; Carbon N. Y. 1998, 36, 1085. [Crossref]

47 Yang, T.; Lua, A. C.; J. Colloid Interface Sci. 2003, 267, 408. [Crossref]

48 Zhang, T.; Walawender, W. P.; Fan, L. T.; Fan, M.; Daugaard, D.; Brown, R. C.; Chem. Eng. J. 2004, 105, 53. [Crossref]

49 Lillo-Ródenas, M. A.; Cazorla-Amorós, D.; Linares-Solano, A.; Carbon N. Y. 2003, 41, 267. [Crossref]
-5050 Marsh, H.; Yan, D. S.; O’Grady, T. M.; Wennerberg, A.; Carbon N. Y. 1984, 22, 603. [Crossref]

The experimental apparatus used in this research for the preparation of AC is shown in Figure 1. First, olive stones’ pyrolysis (carbonization) was carried out in a MELLER PS205 horizontal furnace (New Hampshire, USA) at 400 and 600 °C for 1 h with a heating rate of 5 °C min-1 and a N2 (Linde, 99.995%, Santiago, Chile) flow rate of 150 mL min 11 Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref]. About 40 g of raw material in the alumina ceramic crucible was placed into the furnace. After pyrolysis, the furnace was cooled to room temperature (ca. 20 °C) with a N2 flow. The designations OS-400 and OS 600 refer to olive stones that have been pyrolyzed at 400 and 600 °C, respectively. The pyrolytic char produced at 600 °C was used for ACs prepared in a two-step procedure.

Figure 1
Schematic diagram of the pyrolysis/activation experimental system.

CO2 activation

The pyrolytic char (4 g) was activated at 850 °C (heating rate of 10 °C min-1) with a CO2 (Linde, 99.999%, Santiago, Chile) flow rate of 100 mL min-1, keeping the temperature for 1 h. In this case, the sample was obtained in a two-step activation process. For comparison, the raw material (40 g) was directly activated without being previously pyrolyzed, following the one-step activation procedure. The prepared samples were designated as AC-CO2 1 and AC-CO2 2, and the numbers 1 or 2 correspond to the activation in one or two-step, respectively.

Steam activation

The previously pyrolyzed olive stone (6.5 g) and the raw material (20 g) were activated in the system indicated above at 900 °C for 1 h, keeping a steam/N2 (Linde, 99.99%, Santiago, Chile) ratio equal to 4. Next, the samples were heated to the temperature required for activation in N2 flow (200 mL min-1) with a heating rate of 10 °C min-1. The steam was introduced into the furnace after the temperature was reached. The steam was generated by injecting 0.6 mL min 11 Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref] of liquid water into an evaporator by a peristaltic pump (equivalent to 800 mL min-1 of steam determined at 20 °C and 101.32 kPa). After activation, the samples (AC H2O 1 and AC-H2O 2) were cooled to room temperature under N2 flow.

KOH activation

The pyrolyzed char was impregnated with a concentrated KOH solution in the required proportion to obtain two slurries with a KOH/pyrolyzed char weight ratio of 1 or 2. The slurries were heated at 70 °C until incipient dryness. The samples were then dried overnight at 105 °C. After that, the activation was done for 1 h in a N2 flow of 300 mL min-1 and a heating rate of 5 °C min 11 Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref]. Activation was carried out at 800 and 900 °C. Once activated, the first washing was carried out with 0.1 mol L-1 HCl (Merck, 37%, Darmstadt, Germany) solution and later with distilled water until obtaining a pH > 5. The samples were identified as AC-KOH 1-800, AC KOH 1-900, AC-KOH 2-800, and AC-KOH 2-900, whose numbers 1 or 2 are related to KOH/pyrolyzed weight ratio, and the number 800 or 900 corresponds to the activation temperature. The one step method was not evaluated for this activation method because the previous test resulted in a low yield (< 11%). It is desirable to produce an AC that has a relatively high yield.

H3PO4 activation

The olive stone and pyrolytic char were mixed with an adequate quantity of a H3PO4 solution to obtain a H3PO4/sample weight ratio of 1.75. This mixture was stirred at 70 °C until dry and kept on a stove at 105 °C overnight. It was then activated for 1 h at 600 °C (heating rate: 3 °C min 11 Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref]; N2 flow: 300 mL min-1). Finally, the carbon was washed with hot water until pH > 5 and dried at 105 °C to be later stored. The ACs were named AC-H3PO4 1 and AC-H3PO4 2, where the numbers 1 and 2 indicate if the sample was activated in one or two-step.

On completion, the resultant ACs by different methods were weighed to determine their yield. Here, the global yield (YG) of char or AC is calculated by dividing the sample weight after activation (Wf) by the weight of the raw material (Wi) using the following equation:

(1) Y G ( % ) = ( W f / W i ) × 100

Using the pyrolyzed sample as the starting point, an equation was used to figure out the yield (Yp) of the ACs prepared with the two-step method:

(2) Y p ( % ) = ( W p / W i ) × 100

where Wp is the weight of pyrolyzed sample.

Characterization of ACs

Proximate analysis was done following the ASTM D15065151 ASTM D1506: Standard Test Methods or Carbon Black - Ash Content, West Conshohocken, 2020. and ISO 5625252 ISO 562: Hard Coal and Coke-Determination of Volatile Matter, Switzerland, 2010. standards to determine the content of ash and volatiles, respectively, while the moisture was determined after drying the sample at 105 °C overnight. Ultimate analysis was performed in a Leco CHNS 628 (Michigan, USA) elemental analyzer equipment.

Textural properties were obtained from the N2 adsorption-desorption isotherm at -196 °C with a Micromeritics 3-Flex instrument (Georgia, USA). The samples (30-50 mg) were pretreated at 300 °C in a vacuum condition for 4 h in the Micromeritics SmartVacPrep equipment (Georgia, USA). The Brunauer-Emmett-Teller (BET) method was used to determine the specific surface area, employing the adsorption branch of the isotherms in the range of 0.02 ≤ P/P0 ≤ 0.25.5353 Leofanti, G.; Padovan, M.; Tozzola, G.; Venturelli, B.; Catal. Today 1998, 41, 207. [Crossref],5454 Blanco, E.; Sepulveda, C.; Cruces, K.; García-Fierro, J. L.; Ghampson, I. T.; Escalona, N.; Catal. Today 2020, 356, 376. [Crossref] The amount of N2 adsorbed at P/P0 = 0.99 corresponds to the total pore volume, including the micro and mesopores.5353 Leofanti, G.; Padovan, M.; Tozzola, G.; Venturelli, B.; Catal. Today 1998, 41, 207. [Crossref] All pores in the adsorbent are assumed to be filled with condensed gas at this point. The Dubinin-Radushkevich equation was used to calculate the volume of micropores.5555 Gil, A.; Grange, P.; Colloids Surf., A 1996, 113, 39. [Crossref] The volume of mesopores was obtained as the difference between the total volume and the volume of micropores.

The sodium thiosulfate volumetric method was used to determine the iodine number expressed as the milligrams of iodine adsorbed by 1 g of carbon.5656 ASTM-D4607-14: Standard Test Method for Determination of Iodine Number of Activated Carbon, West Conshohocken, 2014. Briefly, the powder AC and a 0.05 mol L-1 iodine solution (Supelco, Reag. Ph Eur, Darmstadt, Germany) are mixed, shaken, and then the filtrate is submitted to titration against a standardized 0.1 mol L-1 Na2S2O3 solution (Merck, Darmstadt, Germany).

The surface oxygenated functional groups were determined following the Boehm method.5757 Boehm, H. P.; Carbon N. Y. 1994, 32, 759. [Crossref] The four 0.1 g of samples were mixed with 20 mL of 0.05 mol L-1 of NaOH (Winkler, 98.6%, Santiago, Chile), Na2CO3 (Sigma-Aldrich, 99.5%, Massachusetts, USA), NaHCO3 (Winkler, > 99%, Santiago, Chile), and HCl (Merck, 37%, Darmstadt, Germany) solutions, respectively. The vials were sealed and stirred for 24 h in a thermal bath at 25 °C. The suspensions are filtered and titrated by 0.05 mol L-1 HCl or 0.05 mol L-1 NaOH solutions. The numbers of the acidic sites were calculated considering that NaOH neutralizes carboxylic, lactonic and phenolic groups; Na2CO3 neutralizes carboxylic and lactonic, and NaHCO3 neutralizes only carboxylic groups. The amount of HCl that reacted with the AC was used to calculate the number of basic sites.

The chemical nature of the surface was analyzed by temperature-programmed desorption followed by mass spectrometry (TPD-MS), using the same apparatus described in the textural properties test (Micromeritics 3-Flex instrument) coupled with a mass spectrometer (Cirrus 2, MKS Spectra Product, Andover-Massachusetts, USA). The sample (0.05 g) was pre-conditioned in He flow at 45 °C for 1 h before the desorption runs. After that, the sample was heated up to 1000 °C at a heating rate of 10 °C min-1 in He flow (100 mL min-1). A quadrupole mass spectrometer was used to detect the amount of CO2 and CO desorbed as a function of temperature. Calibration was done with calcium oxalate monohydrate for quantification.

The pH of the point of zero charge (pHpzc) is the pH at which the surface’s net charge is zero.5858 Benzekri, M. B.; Benderdouche, N.; Bestani, B.; Douara, N.; Duclaux, L.; J. Mater. Environ. Sci. 2018, 9, 272. [Crossref] 0.1 g of samples were placed in Erlenmeyer flasks with 50 mL of NaCl (Riedel-de Haën, 99.8%, Germany) 0.01 mol L-1 solution. The initial pH (pHi) was adjusted between 2-12 values using NaOH or HCL solutions of 0.1 mol L-1 and a PL-700PC Gondo instrument pH meter (Taipei, Taiwan). Then, the suspension was shaken for 48 h, and the pH of the filtrate was measured (pHf). The pHpzc was obtained from the plot of pHf vs. pHi and corresponded to the value where pHf is equal to pHi.

The surface functional groups of carbons were identified using the Fourier transform infrared spectroscopy (FTIR) technique. A Bruker Alpha FT-IR spectrometer was used to record the samples’ FTIR spectra between 4000 and 400 cm-1 (Billerica, Massachusetts, USA). First, pellets were prepared by mixing carbon and KBr (1:400 carbon/KBr weight ratio) in an agate mortar. Then, in a hand press, the resulting mixture was compressed.

NH3 adsorption

Adsorption equilibrium data of NH3 were carried out in the same equipment employed for the textural properties test (Micromeritics 3-Flex instrument). Before measurement, 50 mg of sample was treated in He flow at 300 °C for 4 h and then exposed to the increasing pressure of anhydrous NH3 (Indura, 99.5%, Santiago, Chile) up to 90 kPa at room temperature. The adsorption equilibrium amount was calculated at the terminal pressure from the final adsorption amount.

The experimental data were fitted to the Langmuir and Freundlich isotherms.5959 Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 345, 402. [Crossref] The Langmuir isotherm is expressed as:

(3) q c = ( a m × b × P ) / ( 1 + b × P )

where qc is the adsorbed NH3 amount per unit mass of solid (cm3 g-1), am is the maximum adsorption capacity of the solid (mg g-1), b is the Langmuir adsorption constant related to the adsorption energy (kPa-1), and P is the NH3 pressure (kPa).

The Freundlich isotherm is given by:

(4) q c = K F × P 1 / n

where KF (cm3 g-1 kPa-1/n) and n are the constants, relate to the relative capacity adsorption of adsorbent and the adsorption intensity, respectively.5959 Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 345, 402. [Crossref],6060 Djelloul, C.; Hamdaoui, O.; Desalin. Water Treat. 2015, 56, 2966. [Crossref]

Nonlinear regression is used to determine all the parameters from the experimental points. The fitness degree was estimated by the coefficient of determination (R2), and the average relative error (ARE) was calculated by the equation:6161 Hashem, A.; Al-Kheraije, K. A.; J. Environ. Prot. 2013, 4, 280. [Crossref]

(5) ARE ( % ) = ( 1 N q exp q c q exp / N ) × 100

where qexp is the experimentally adsorbed amount of NH3, qc is the amount of NH3 calculated using the models, and N is the number of adsorption points.

Results and Discussion

Starting and pyrolyzed material

Table 1 shows the properties of the raw and pyrolyzed olive stone. The olive stone is characterized by the high volatiles content and carbon and low ash values.6262 Mediavilla, I.; Barro, R.; Borjabad, E.; Peña, D.; Fernández, M. J.; Renewable Energy 2020, 160, 374. [Crossref],6363 Bartocci, P.; D’Amico, M.; Moriconi, N.; Bidini, G.; Fantozzi, F.; Energy Procedia 2015, 82, 374. [Crossref] For comparison purposes, the olive stone was pyrolyzed at two temperatures (400 and 600 °C). This sample’s proximate and ultimate analysis results are compared with those reported by the references mentioned above, although the moisture can vary between 5 to 10%.6464 Moubarik, A.; Grimi, N.; Food Res. Int. 2015, 73, 169. [Crossref] Sulfurous was not detected in the sample. After pyrolysis of olive stone, the content of volatiles decreased while the carbon increased. It is well known that the increase in temperature promotes the degradation of the constituents of the olive stone (hemicellulose, cellulose and lignin), and these components can be volatilized, and inorganic material (ash) is concentrated.6565 Bedoui, A.; Souissi-Najar, S.; Ouederni, A.; Int. J. Innovation Appl. Stud. 2016, 18, 938. [Crossref] As a result of the large quantity of volatiles generated during pyrolysis, yields diminish as the temperature rises from 400 to 600 °C.1111 Lua, A. C.; Yang, T.; Guo, J.; J. Anal. Appl. Pyrolysis 2004, 72, 279. [Crossref] The OS-600 sample was used to produce the AC due to the higher C content compared to OS-400.

Table 1
Properties of raw and pyrolyzed olive stone

Characterization of ACs

The proximate and ultimate analysis results and the yield of the ACs are given in Table 2. The values are similar for all samples and close to the commercial ACs analyzed. The activation increases the carbon content, and the ash tends to concentrate. The %H decreased after the activation treatments, and the nitrogen content remained low, varying between 0.5-1%.3434 Moreno-Castilla, C.; Carrasco-Marín, F.; López-Ramón, M. V.; Alvarez-Merino, M. A.; Carbon N. Y. 2001, 39, 1415. [Crossref]

Table 2
Proximate and ultimate analysis and yield of ACs

It is desirable to produce an AC with a relatively high yield, and depending on whether activation was carried out in one- or two-step, this parameter will be affected. When the global yield is considered, the one- and two-step activations report lower yields (20%), except when H3PO4 is used as an activating agent for the one-step method (40%). When CO2 is used as an activating agent, and the sample is activated in one-step (AC-CO2 1), a yield of 17% was obtained. The global yield remained in the same order (19%) when the activation was carried out in two steps (AC-CO2 2 sample). The yield (Yp) significantly increased up to 70% when calculated from the pyrolyzed olive stone as a starting material. It has been reported that the partial gasification of the raw material or char exposed to CO2 occurs by the following equation:88 Román, S.; González, J. F.; González-García, C. M.; Zamora, F.; Fuel Process. Technol. 2008, 89, 715. [Crossref]

(6) C + C O 2 2 C O

In this sense, both the effect of a high temperature and CO2 presence promotes the precursor decomposition and increases the extent of the reaction,55 Yang, K.; Peng, J.; Xia, H.; Zhang, L.; Srinivasakannan, C.; Guo, S.; J. Taiwan Inst. Chem. Eng. 2010, 41, 367. [Crossref] causing a decrease in yield (or increased burn-off). In the case of two-step activation, the yield is higher since a large part of the volatiles have already been released.

In steam activation, water reacts with the carbon of the sample to produce carbon monoxide and hydrogen, according to:88 Román, S.; González, J. F.; González-García, C. M.; Zamora, F.; Fuel Process. Technol. 2008, 89, 715. [Crossref]

(7) C + H 2 O C O + H 2

Therefore, the carbon is consumed, and porosity (micro and mesopores) can be generated. The use of steam or CO2 makes physical activation more advantageous than chemical activation, requiring more expensive and toxic chemicals and consecutive washing processes to eliminate excess reagents. The global yield obtained for steam activation in one- or two-step was low, registering values between 10 and 11%. During this process, weight loss occurs both due to the evolution of volatile material and the elimination of carbon atoms through steam reaction (equation 6). This low yield agrees with previous research,6666 Minkova, V.; Razvigorova, M.; Bjornbom, E.; Zanzi, R.; Budinova, T.; Petrov, N.; Fuel Process. Technol. 2001, 70, 53. [Crossref] which indicated that low solid yields (or high burn-off) and high generation of volatiles are associated with steam’s capacity to penetrate the solid material, helping the removal and efficient desorption of volatile products. In the case of the AC-H2O 2 sample, a yield of 37% was obtained, employing the pyrolyzed sample mass as starting material. This result is because the sample activated with steam was previously pyrolyzed, and a large part of the volatiles was removed.

For chemical activation, specifically the KOH activation, it has been established that this method consists of redox reactions, with the carbon oxidation to CO or CO2.4949 Lillo-Ródenas, M. A.; Cazorla-Amorós, D.; Linares-Solano, A.; Carbon N. Y. 2003, 41, 267. [Crossref] As shown below, KOH activation involves a sequence of reactions, including dehydration, water gas reaction, and reduction by hydrogen and carbon.6767 Gebreegziabher, T. B.; Wang, S.; Nam, H.; J. Environ. Chem. Eng. 2019, 7, 103234. [Crossref] During these reactions, the biomass is broken down, resulting in the generation of micropores and mesopores structures.

(8) 2 K O H K 2 O + H 2 O ( removal of water )
(9) C + H 2 O H 2 + C O ( water-gas reaction )
(10) C O + H 2 O H 2 + C O 2 (water-gas shift reaction)
(11) K 2 O + H 2 K + H 2 O (reduction by hydrogen)
(12) K 2 O + C 2 K + C O (reduction by carbon)

The increase in the KOH/precursor ratio and the temperature causes a slight decrease in the global yield, obtaining 20 and 17% of yield for AC-KOH 1-800 and AC KOH 2-900 samples, respectively. These yields were 76 and 62% for the previously mentioned samples when calculated from the pyrolyzed material mass, and the effect of the KOH/precursor ratio and temperature on the yield was more evident. Metallic potassium is obtained due to KOH reduction or its transformation products (e.g., K2CO3).5050 Marsh, H.; Yan, D. S.; O’Grady, T. M.; Wennerberg, A.; Carbon N. Y. 1984, 22, 603. [Crossref] The solid particle breaks down when potassium is intercalated within the graphitic lamellar structure, making granular carbon preparation difficult. This explains why the carbon obtained from this activation turned into a fine powder, and high temperatures or KOH concentration affected the yield.

In addition, the direct activation of the raw olive stone produces a very light material due to the breakdown of the precursor particles after KOH solution impregnation. The attack of hydroxyl ions causes the fragmentation and dissolution of lignin and hemicellulose.3939 Molina-Sabio, M.; Rodríguez-Reinoso, F.; Colloids Surf., A 2004, 241, 15. [Crossref] It explains the meager yield obtained (< 11%) in the preliminary test using unpyrolyzed olive stone and why it was not used in the AC preparation.

When H3PO4 was used as an activating agent, a YG of 40 was obtained for the AC-H3PO4 1 sample, while the AC-H3PO4 2 sample registered a YG of 18% (68%, considering the pyrolyzed material). This trend is similar to that observed for the evaluated physical activations. As mentioned before, the sample impregnation generates reactions that promote cellulose fragmentation and other structural components of the olive stone, so the raw material is more affected than the pyrolyzed char. Because the acid separates the cellulose fibers and causes partial depolymerization of hemicellulose and lignin, the particle treated with H3PO4 becomes elastic, causing a decrease in mechanical resistance and particle swelling. Similarly, because many tars have been detected on the particle’s surface, impregnation promotes carbon conversion.4545 Solum, M. S.; Pugmire, R. J.; Jagtoyen, M.; Derbyshire, F.; Carbon N. Y. 1995, 33, 1247. [Crossref] These tars come from acid-catalyzed cellulose depolymerization, dehydration, and condensation, resulting in aromatic and reactive compounds with some cross-linking.

Figure 2 depicts the nitrogen adsorption-desorption isotherms for the physically and chemically activated samples, while Table 3 reports the textural properties obtained from the adsorption isotherms. All the carbons presented type I isotherms, characteristic of predominant microporous solids. For samples activated with CO2 and KOH, a percentage of micropores (%Vmi/VT) higher than 90% was obtained, while the carbons obtained by steam and H3PO4 activation showed a micro-mesoporous structure.88 Román, S.; González, J. F.; González-García, C. M.; Zamora, F.; Fuel Process. Technol. 2008, 89, 715. [Crossref],3838 Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A. F.; Maldonado-Hódar, F. J.; Carrasco-Marín, F.; Electrochim. Acta 2017, 229, 219. [Crossref],6767 Gebreegziabher, T. B.; Wang, S.; Nam, H.; J. Environ. Chem. Eng. 2019, 7, 103234. [Crossref],6868 Hu, Z.; Srinivasan, M. P.; Microporous Mesoporous Mater. 2001, 43, 267. [Crossref]

Table 3
Textural properties of samples prepared by physical and chemical activation

Figure 2
N2 adsorption-desorption isotherms at -196 °C of ACs. Closed symbols: adsorption curve and open symbols: desorption curve.

For micro-mesoporous carbons, an H4-type hysteresis loop was observed, usually found in solids that consist of aggregates or agglomerates of particles with slit-shaped pores.33 Chang, C.-F.; Chang, C.-Y.; Tsai, W.-T.; J. Colloid Interface Sci. 2000, 232, 45. [Crossref] These samples showed percentages of micropores between 60-78%, with an average pore diameter very close to 2 nm, indicative that small mesopores are present. It has been reported that CO2 activation favors the creation of microporosity while the use of steam increases microporosity at the beginning of the activation, which are then transformed into mesopores, obtaining carbons with a lower percentage micropore than CO2-activated samples.3535 Rodríguez-Reinoso, F.; Molina-Sabio, M.; González, M. T.; Carbon N. Y. 1995, 33, 15. [Crossref],6969 Molina-Sabio, M.; González, M. T.; Rodriguez-Reinoso, F.; Sepúlveda-Escribano, A.; Carbon N. Y. 1996, 34, 505. [Crossref]

The chemically activated samples showed the highest surface areas, highlighting the AC-KOH 2-900 (1823 m2 g-1) and AC-H3PO4 1 (1986 m2 g-1) samples. For the KOH-activated samples, increasing the KOH/precursor ratio (from 1 to 2) and the activation temperature (from 800 to 900 °C) improves the surface area. This result can be expected because more KOH is intercalated into the material structure, favoring the porosity generation during impregnation and subsequent activation.3838 Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A. F.; Maldonado-Hódar, F. J.; Carrasco-Marín, F.; Electrochim. Acta 2017, 229, 219. [Crossref],7070 El-Hendawi, A.; Appl. Surf. Sci. 2009, 255, 3723. [Crossref] It has been reported the presence of mesopores when wood carbons were activated with H3PO4, while the carbons obtained by KOH activation reported a microporous structure.7171 Salame, I. I.; Bandosz, T. J.; Ind. Eng. Chem. Res. 2000, 39, 301. [Crossref]

The one-step sample activated with CO2 (AC-CO2 1) has a surface area of 1011 m2 g-1, a higher value than the sample activated in a two-step (AC-CO2 2, 810 m2 g-1). However, the percentage of micropores did not change (ca. 90%). This same behavior was reported for H3PO4-activated carbons, where the surface area of AC-H3PO4 1 was higher than AC-H3PO4 2. The opposite trend was observed for the steam-activated sample, whose surface area of the AC-H2O 1 sample (597 m2 g-1) was lower than that activated in a two-step (AC-H2O 2, 1511 m2 g-1). These results indicate the importance of the initial contact surface with the activating agent and the temperature, activation time, activating agent flow, and heating ramp that influence the diffusion and removal of gaseous products from the surface of the coal.44 Rodríguez-Reinoso, F.; Molina-Sabio, M.; Carbon N. Y. 1992, 30, 1111. [Crossref],55 Yang, K.; Peng, J.; Xia, H.; Zhang, L.; Srinivasakannan, C.; Guo, S.; J. Taiwan Inst. Chem. Eng. 2010, 41, 367. [Crossref],3434 Moreno-Castilla, C.; Carrasco-Marín, F.; López-Ramón, M. V.; Alvarez-Merino, M. A.; Carbon N. Y. 2001, 39, 1415. [Crossref]

It is well known that the non-carbonized substance interacts with the activating agent more intensely than the carbonized material.3838 Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A. F.; Maldonado-Hódar, F. J.; Carrasco-Marín, F.; Electrochim. Acta 2017, 229, 219. [Crossref] Hence, carbons obtained in one-step activation show higher BET surface area (SBET) than carbons prepared in two-step activation (using OS-600 sample for the steam activation). However, this behavior was not observed for steam-activated samples. It is possible that the one-step activation conditions were very aggressive, consuming the sample and the destruction of pores. This hypothesis is established considering the low global yield obtained for the two activated carbons (AC-H2O 1 and AC-H2O 2) and other investigations,22 Demiral, H.; Demiral, I.; Karabacakoĝlu, B.; Tümsek, F.; Chem. Eng. Res. Des. 2011, 89, 206. [Crossref],33 Chang, C.-F.; Chang, C.-Y.; Tsai, W.-T.; J. Colloid Interface Sci. 2000, 232, 45. [Crossref] indicating that the external burning of carbon particles affects the porosity. For carbons prepared in the two-step activation, the sample carbonized in nitrogen has a more organized structure and less developed porosity than carbonized in steam. This fact is related to a lower reaction rate for the nitrogen-carbonized sample.7272 Savova, D.; Apak, E.; Ekinci, E.; Yardim, F.; Petrov, N.; Budinova, T.; Razvigorova, M.; Minkova, V.; Biomass Bioenergy 2001, 21, 133. [Crossref]

The iodine number allows one to quantify the AC’s adsorptive capacity and can be used to determine pores with dimensions ≥ 1.0 nm.4444 Haimour, N. M.; Emeish, S.; Waste Manage. 2006, 26, 651. [Crossref] Iodine number of the samples with the highest surface area by each method and commercial carbons is reported in Table 4. Chemical KOH activation generates a carbon with high iodine adsorption capacity, with an iodine number of 1030 mg g-1 for AC-KOH 2-900 sample, even higher than those obtained for commercial carbons. Considering that iodine adsorption is indicative of the adsorption capacity in micropores,7373 Sun, K.; Jiang, J. C.; Biomass Bioenergy 2010, 34, 539. [Crossref] this result is well-correlated with its textural properties (Table 3) because a high percentage of micropores was reported (98%). The other carbons showed a lower iodine number than the KOH sample. It may be due to the creation of meso and macropore structures, thus decreasing the adsorptive capacity of iodine.22 Demiral, H.; Demiral, I.; Karabacakoĝlu, B.; Tümsek, F.; Chem. Eng. Res. Des. 2011, 89, 206. [Crossref] Similar values of iodine number for carbons based on olive stone have been reported by other researchers.2929 Martínez, M. L.; Torres, M. M.; Guzmán, C. A.; Maestri, D. M.; Ind. Crops Prod. 2006, 23, 23. [Crossref],3636 Baçaoui, A.; Yaacoubi, A.; Dahbi, A.; Bennouna, C.; Phan Tan Luu, R.; Maldonado-Hodar, F. J.; Rivera-Utrilla, J.; Moreno-Castilla, C.; Carbon N. Y. 2001, 39, 425. [Crossref],7474 Budinova, T.; Ekinci, E.; Yardim, F.; Grimm, A.; Björnbom, E.; Minkova, V.; Goranova, M.; Fuel Process. Technol. 2006, 87, 899. [Crossref]

Table 4
Iodine adsorption and Boehm titration of selected carbons

Boehm titration results of selected carbons are shown in Table 4. The chemistry of the surface and the porosity have a significant impact on the performance of ACs. Acidic groups include carbonyl, carboxyl, phenolic hydroxyl, and lactonic groups, whereas basic groups include pyrene, chromene, quinone, and the graphene structure’s electrons.7575 László, K.; Microporous Mesoporous Mater. 2005, 80, 205. [Crossref] The AC surface’s acid/base nature is determined by the groups and delocalized electrons of the graphitic structure.7676 Radovic, L.; Surface Chemistry of Activated Carbon Materials: State of the Art and Implications for Adsorption, Surfaces of Nanoparticles and Porous Materials; Schwarz, J.; Contescu, C., eds.; Marcel Dekker INC.: New York, 1999. Depending on the activation method and preparation conditions, functional groups on the carbon surface with a wide range of concentrations are favored.7474 Budinova, T.; Ekinci, E.; Yardim, F.; Grimm, A.; Björnbom, E.; Minkova, V.; Goranova, M.; Fuel Process. Technol. 2006, 87, 899. [Crossref]

Acidic surface groups decompose at high temperatures, implying that high activation temperatures favor low acidity.7777 Guo, J.; Lua, A. C.; Mater. Lett. 2002, 55, 334. [Crossref] Because of this, physically activated carbons (like CO2 and steam) tend to be basic, which is in line with the pHpzc values (7.8 and 8.2, respectively) found in Table 4 and shown in Figure S1 (SI section). Therefore, the surface groups in the AC-CO2 1 and AC-H2O 2 samples are mostly basic. The same trend was registered for the AC KOH 2-900 sample. In this sense, carbons prepared from hazelnut bagasse and activated with KOH (KOH/precursor mass ratio between 1-3) have predominantly basic surface groups.66 Demiral, H.; Demiral, I.; Fatma, T.; Karabacako, B.; Surf. Interface Anal. 2008, 40, 616. [Crossref] The sample activated with KOH registered the highest amount of carboxylic groups, with total values of the basic groups slightly higher than acidic groups. This result promotes a little acidic pHpzc value (6.7).

On the other hand, treatment with H3PO4 generates an activated carbon of an acidic nature (pHpzc = 3.2), with fewer carboxylic groups than the sample activated with KOH. However, this sample recorded the highest content of phenolic groups. These results are consistent with those registered in the TPD tests, obtaining a more remarkable CO2 evolution in the sample activated with KOH. In contrast, the one activated with H3PO4 showed a higher evolution of CO.

For AC-H3PO4 1 sample, a higher proportion of acid groups was generated, a result also obtained by other researchers.5858 Benzekri, M. B.; Benderdouche, N.; Bestani, B.; Douara, N.; Duclaux, L.; J. Mater. Environ. Sci. 2018, 9, 272. [Crossref],7474 Budinova, T.; Ekinci, E.; Yardim, F.; Grimm, A.; Björnbom, E.; Minkova, V.; Goranova, M.; Fuel Process. Technol. 2006, 87, 899. [Crossref],7878 Termoul, M.; Bestani, B.; Benderdouche, N.; Belhakem, M.; Adsorpt. Sci. Technol. 2006, 24, 375. [Crossref] In general, the acidic content of chemically prepared ACs was higher than those that were physically activated, and similar pHpzc values were reported.5858 Benzekri, M. B.; Benderdouche, N.; Bestani, B.; Douara, N.; Duclaux, L.; J. Mater. Environ. Sci. 2018, 9, 272. [Crossref],7979 Esteves, B. M.; Morales-Torres, S.; Maldonado-Hódar, F. J.; Madeira, L. M.; Chem. Eng. J. 2021, 411, 128451. [Crossref]

80 Youssef, A.; El-Nabarawy, T.; El-Shafey, E.; Carbon Lett. 2006, 7, 1. [Crossref]
-8181 Şirazi, M.; Aslan, S.; Chem. Eng. Commun. 2021, 208, 1479. [Crossref] The commercial activated carbons Norit GAC 1240 and C Gran showed pHpzc measurements related to their neutral and acid nature.8282 Choi, S.; Choi, W.; Kim, S.; Lee, S.-Y.; Noh, I.; Kim, C. W.; Biomater. Res. 2014, 18, 6. [Crossref]

The total capacity of oxide groups increased using H3PO4 as an activating agent. The generation by surface oxidation of acidic functional groups containing oxygen and/or phosphorous and its attachment to the surface while the porosity is being developed has been proposed.8383 Dastgheib, S. A.; Rockstraw, D. A.; Carbon N. Y. 2001, 39, 1849. [Crossref] Even in the case of ACs prepared using H3PO4, the carboxylic group is thought to be the only strong acidic group detectable by NaHCO3.6969 Molina-Sabio, M.; González, M. T.; Rodriguez-Reinoso, F.; Sepúlveda-Escribano, A.; Carbon N. Y. 1996, 34, 505. [Crossref],8484 Toles, C. A.; Marshall, W. E.; Johns, M. M.; Carbon N. Y. 1999, 37, 1207. [Crossref],8585 Mahmood, T.; Ali, R.; Naeem, A.; Hamayun, M.; Aslam, M.; Process Saf. Environ. Prot. 2017, 109, 548. [Crossref] The -OH group linked to phosphorus-containing acids adsorbed on the surface has been partly responsible for the surface’s strong acidity (i.e., -PO2H2).8686 Benaddi, H.; Bandosz, T. J.; Jagiello, J.; Schwarz, J. A.; Rouzaud, J. N.; Carbon N. Y. 2000, 38, 669. [Crossref] Based on the investigation mentioned before, it is suggested that the acidic groups detected should be considered as both oxygen-containing and oxygen/phosphorus acidic groups.

The surface chemistry of the ACs was evaluated using temperature-programmed desorption (TPD) experiments coupled to mass. The temperature at which CO and CO2 are evolved during TPD can characterize the functional groups that contain oxygen. The CO2 and CO desorption profiles obtained by TPD-MS are shown in Figure 3, while the amounts desorbed are compiled in Table 5. It has been proposed as a general trend that CO2 evolves at temperatures between 300 and 800 °C and corresponds to the degradation of the surface oxygen groups that are more acidic and less stable. At temperatures near 300 °C, carboxylic groups degrade to CO2, followed by lactones and anhydrides (which give both CO and CO2) at temperatures above 800 °C. CO is produced by the decomposition of the oxygen groups that are the most stable and least acidic, such as anhydride and phenolic groups at 600 °C, and of carbonyls and quinone at temperatures that can reach 900 °C.8787 Gonçalves, M.; Sánchez-García, L.; de Oliveira Jardim, E.; Silvestre-Albero, J.; Rodríguez-Reinoso, F.; Environ. Sci. Technol. 2011, 45, 10605. [Crossref]

88 Figueiredo, J. L.; Pereira, M. F. R.; Freitas, M. M. A.; Órfão, J. J. M.; Carbon N. Y. 1999, 37, 1379. [Crossref]

89 Figueiredo, J. L.; Pereira, M. F. R.; Catal. Today 2010, 150, 2. [Crossref]

90 Ro, K. S.; Lima, I. M.; Reddy, G. B.; Jackson, M. A.; Gao, B.; Agriculture 2015, 5, 991. [Crossref]
-9191 Zhou, J. H.; Sui, Z.-J.; Zhu, J.; Li, P.; Chen, D.; Dai, Y.-C.; Yuan, W.-K.; Carbon N. Y. 2007, 45, 785. [Crossref]

Table 5
Amounts of CO and CO2 desorbed measured by TPD-MS

Figure 3
Desorption profiles of ACs obtained by TPD-MS. (a): CO2 evolution, (b): CO evolution.

AC-KOH 2-900 sample (Figure 3a) showed the highest value for CO2 evolution (3.306 mmol g-1), which can be explained by the high degree of more acidic surface groups, such as carboxylic groups. The presence of this acidic group was also corroborated by the Boehm and FTIR characterization. As will be discussed later, the existence of this surface oxygen group plays an essential role in NH3 adsorption. For AC-H3PO4 1 sample, the CO2 TPD profile shows the majority presence of lactones and/or anhydrides groups. At the same time, the region related to the decomposition of carboxylic groups did not have a significant contribution to the profile. This observation could corroborate the hypothesis raised concerning the overestimation of carboxylic groups in the Boehm treatment.

The CO desorption profiles are shown in Figure 3b. AC H3PO4 1 sample had the highest CO evolution (5.170 mmol g-1), registering broadband between 350 and 1000 °C with a maximum at 848 °C and a shoulder near 686 °C. The signal at 686 °C is attributed to phenol and hydroquinones, while the maximum at 848 °C is assigned to carbonyl and quinones.9292 Domingo-García, M.; Groszek, A. J.; López-Garzón, F. J.; Pérez-Mendoza, M.; Appl. Catal., A 2002, 233, 141. [Crossref]

The difference in the oxygenated surface groups was noticed depending on the chemical agent used for the activation. For KOH activated sample (AC-KOH 2-900), oxygenated surface groups degrade primarily as CO2 (anhydrides, lactones, carboxylic acid, etc.) during TPD tests, while for H3PO4 activated sample (AC-H3PO4 2), those groups decompose as CO (carbonyl, semiquinone, etc.) are the most numerous.3838 Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A. F.; Maldonado-Hódar, F. J.; Carrasco-Marín, F.; Electrochim. Acta 2017, 229, 219. [Crossref] In this sense, the oxygen-phosphorous bond in the C-O-P system is the weakest in H3PO4 activated samples. Therefore, this bond is broken at high temperatures, leaving only an O atom linked to a carbon site and favoring the CO desorption.9393 Wu, X.; Radovic, L. R.; Carbon N. Y. 2006, 44, 141. [Crossref]

FTIR spectra of the obtained ACs are shown in Figure 4. The assignments of various bands observed are listed in Table 6. The presence of the oxygen functional groups was determined by the FTIR, which agrees with the Boehm results. The samples’ spectra were very similar, and an intense band at 3428 cm-1 for AC-H2O 2 sample can be noticed. Considering that this sample was activated using steam and the band can also be assigned to adsorbed water,9494 Song, J.; Zou, W.; Bian, Y.; Su, F.; Han, R.; Desalination 2011, 265, 119. [Crossref] the contribution of water on the spectrum is not ruled out.

Table 6
Assignments of infrared bands observed in the ACs prepared

Figure 4
FTIR (KBr) spectra of ACs: (a) AC-CO2 1; (b) AC-H2O 2; (c) AC KOH 2-900; (d) AC-H3PO4 1; (e) Norit C Gran.

It has been reported in H3PO4 ACs prepared from woody biomass, a broadband between 900-1300 cm-1 with a maximum at 1116-1142 cm-1. This band is related to oxygen and phosphorous compounds in the ACs.7474 Budinova, T.; Ekinci, E.; Yardim, F.; Grimm, A.; Björnbom, E.; Minkova, V.; Goranova, M.; Fuel Process. Technol. 2006, 87, 899. [Crossref] In our results, considering the similarity of the carbons’ spectra that were prepared using different methods and the overlap of the absorption bands, an unambiguous assignment is complex, and just the presence of phosphorous compounds can be suggested.

For Norit C Gran commercial AC, the intense band at 1582 cm-1 indicates a more non-polar (hydrophobic) surface than the other prepared carbons.1515 Nagalakshmi, T. V.; Emmanuel, K. A.; Suresh Babu, C.; Chakrapani, C.; Divakar, P. P.; Int. Lett. Chem., Phys. Astron. 2015, 54, 189. [Crossref]

NH3 adsorption

NH3 adsorption isotherms of the selected samples at 20 °C are plotted in Figure 5. The maximum amount of NH3 adsorbed was determined from these isotherms and is reported in Table 7. The volume adsorbed indicates a high adsorption capacity for all the ACs prepared compared to other carbons discussed in the literature and the commercial ones used as references. For example, AC-KOH 2-900 AC reports the higher NH3 volume adsorbed (332.9 cm3 g-1), corresponding to 252 mg g-1. The high adsorption capacity obtained for the samples could be due to the oxygen functional groups, especially the carboxylic groups.100100 Asada, T.; Ohkubo, T.; Kawata, K.; Oikawa, K.; J. Heal. Sci. 2006, 52, 585. [Crossref]Figure 6 depicts the interaction between the ammonia molecule and the carboxylic groups of ACs.

Table 7
NH3 adsorption capacity of some adsorbents

Figure 5
NH3 adsorption isotherms at 20 °C for selected samples. Symbols: experimental data, solid line: Freundlich model (a), dash line: Langmuir model (b).

Figure 6
Interaction of ammonia with the carboxylic groups of AC.

Boehm titration, FTIR and TPD results have demonstrated the presence of acidic functional groups (phenolic, lactonic and carboxylic) that can explain the high NH3 adsorption capacity obtained. It has been found that an increase in the number of functional oxygen groups on the surface of modified carbons favors NH3 adsorption capacity and is related to the acid site type present. Due to the electronegative nature of nitrogen, the hydrogen atoms of the NH3 molecule interact with the oxygen atoms linked as hydroxyl (H-O) and carbonyl (C-O) groups on the surface because of the electrostatic forces (see Figure 6).105105 Shan, X.-m.; Zhu, S.-q.; Zhang, W.-h.; J. China Univ. Min. Technol. 2008, 18, 261. [Crossref]

In this sense, TPD results of the AC-KOH 2-900 sample indicated the presence of a higher proportion of more acidic and less stable surface groups than other samples and explained the high NH3 adsorption capacity obtained.9090 Ro, K. S.; Lima, I. M.; Reddy, G. B.; Jackson, M. A.; Gao, B.; Agriculture 2015, 5, 991. [Crossref] This assertion is established by correlating the adsorption capacity of NH3 with the amount of more acidic surface groups (CO2) that was released during the TPD test (Figure 7a). A linear trend was observed (R2 = 0.857), although the increase in the NH3 adsorption capacity with the more acidic surface oxygen groups was clearer for the AC-KOH 2-900 sample. In this sense, Gonçalves et al.8787 Gonçalves, M.; Sánchez-García, L.; de Oliveira Jardim, E.; Silvestre-Albero, J.; Rodríguez-Reinoso, F.; Environ. Sci. Technol. 2011, 45, 10605. [Crossref] showed a linear correlation between the quantity of the more acidic oxygen surface groups and the NH3 adsorption capacity in a commercial carbon (MAST).

Figure 7
Relationship between NH3 adsorption capacity, more acidic surface groups (a), and micropores volume (b).

In order to evaluate the influence of microporosity on the adsorption capacity of NH3, these two parameters were plotted and are shown in Figure 7b. There is no relationship between the parameters that can be corroborated with the behavior of AC-KOH 2-900 and AC-H3PO4 1 samples with an equal volume of micropores (ca. 0.73 cm3 g-1) but very different NH3 adsorption (252 vs. 156 mg g-1, respectively). This tendency is identical if it is plotted the NH3 adsorption against the surface area and mesopore volume. This behavior was also reported by Yeom et al.102102 Yeom, C.; Kim, Y.; Environ. Eng. Res. 2017, 22, 401. [Crossref] in a commercial AC, who indicated that the NH3 adsorption capacity was unaffected by the surface area. However, other authors have established the preference for adsorbents with large surface areas and small pore sizes for NH3 removal.101101 Helminen, J.; Helenius, J.; Paatero, E.; Turunen, I.; J. Chem. Eng. Data 2001, 46, 391. [Crossref],106106 Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 348, 615. [Crossref] Additionally, mesoporous samples have been proposed to adsorb ammonia because ordered and regular porous structures with well-defined channels favor the adsorption of NH3.5959 Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 345, 402. [Crossref],102102 Yeom, C.; Kim, Y.; Environ. Eng. Res. 2017, 22, 401. [Crossref] These different results suggest that the nature and preparation conditions of the sample influence the NH3 adsorption. It could be improved by the surface area, porosity, and/or surface chemistry. Thus, from our results, the good adsorption behavior obtained for the KOH-activated sample could be explained by the combination of the high microporosity and adequate surface acidity. The activation with KOH effectively promoted the creation of oxygenated surface groups (carboxylic groups) and high microporosity (98%) with the best adsorption properties to remove NH3. The results obtained in this study indicated the best NH3 adsorption for carbons chemically activated following the order: AC-KOH 2-900 > AC H3PO4 1 > AC-CO2 1 > AC-H2O 2.

Regarding the adsorption models, the experimental data were adjusted to the Langmuir and Freundlich adsorption isotherm models. The Langmuir model implies monolayer adsorption on a homogeneous surface, with all active sites having the same energy and being equivalent. In addition, the Langmuir model considers dynamic equilibrium and no adsorbate interaction.107107 Sangon, S.; Hunt, A. J.; Attard, T. M.; Mengchang, P.; Ngernyen, Y.; Supanchaiyamat, N.; J. Cleaner Prod. 2018, 172, 1128. [Crossref] The Freundlich model is based on the following assumptions: (i) multilayer adsorption on heterogeneous surfaces, (ii) adsorbate interactions and (iii) increase in the adsorption capacity with the concentration of the analyte.1414 Luo, L.; Wu, X.; Li, Z.; Zhou, Y.; Chen, T.; Fan, M.; Zhao, W.; R. Soc. Open Sci. 2019, 6, 190523. [Crossref]

According to the results in Table 8, the lower ARE and the closest R2 values to 1 indicate a good correlation for the Freundlich adsorption isotherm related to multilayer adsorption on a heterogeneous surface.1414 Luo, L.; Wu, X.; Li, Z.; Zhou, Y.; Chen, T.; Fan, M.; Zhao, W.; R. Soc. Open Sci. 2019, 6, 190523. [Crossref] Thus, the Freundlich model best describes NH3 adsorption on prepared and reference ACs. In this sense, several authors have reported the same behavior in the fitting of the experimental data to the Freundlich model,5959 Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 345, 402. [Crossref],108108 Liang, P.; Yu, H.; Huang, J.; Zhang, Y.; Cao, H.; MATEC Web Conf. 2016, 67, article 07006. [Crossref],109109 Hayhurst, D. T.; Chem. Eng. Commun. 1980, 4, 729. [Crossref] and the value of n > 1 indicates favorable adsorption conditions.6767 Gebreegziabher, T. B.; Wang, S.; Nam, H.; J. Environ. Chem. Eng. 2019, 7, 103234. [Crossref]

Table 8
Fit of the experimental data of NH3 adsorption at 20 °C to the Langmuir and Freundlich isotherms

Conclusions

In this study, ACs using olive stones were prepared. Depending on the activation method, the surface area of carbons ranges between 1000-1986 m2 g-1, indicating that olive stone was successfully activated. Type I isotherms were observed in all the samples, characteristics of microporous solids; however, the isotherms obtained for steam and H3PO4 activated samples indicated the contribution of mesopores. KOH-AC reported the highest iodine adsorptive capacity (1030 mg g-1), which correlates with the high percentage of microporous obtained (98%). Surface area results showed the importance of the initial contact surface with the activating agent depending on whether the activation is in one or two steps, affecting the diffusion and removal of gaseous products from the sample. The presence of acidic functional groups (carboxylic, lactonic, phenolic) was observed by Boehm titration, FTIR, and TPD characterizations. It can explain the NH3 adsorption capacity obtained, especially for AC KOH 2-900 sample (252 mg NH3 g carbon-1). The TPD profile of KOH AC showed the highest CO2 evolution because more acidic surface groups were reported (carboxylic groups), and a linear trend was observed between the amount of NH3 adsorbed and these surface groups. However, the surface area and micropore volume of samples did not affect the adsorption capacity of NH3. The NH3 adsorption amounts of the physical and chemical prepared ACs are in the following order: AC-KOH 2-900 > AC-H3PO4 1 > AC-CO2 1 > AC-H2O 2. The good performance for the NH3 adsorption obtained for the KOH-activated sample could be explained by the combination of the high microporosity and adequate surface acidity. The Freundlich model best fit the experimental data, implying that the multilayer adsorption on a heterogeneous surface was dominant.

Supplementary Information

Supplementary information (commercial ACs, Figure S1) is available free of charge at http://jbcs.sbq.org.br as PDF file.

Acknowledgments

This work was supported by the Fundación para la Innovación Agraria FIA-Project PYT-2019-0158. N.E. acknowledges Fondequip No. EQM160070 and the Millennium Science Initiative Program NCN17_040. V.F. acknowledges ANID BASAL (grant number ACE210012 and FB 210015).

References

  • 1
    Saleem, J.; Shahid, U. B.; Hijab, M.; Mackey, H.; McKay, G.; Biomass Convers. Biorefin. 2019, 9, 775. [Crossref]
  • 2
    Demiral, H.; Demiral, I.; Karabacakoĝlu, B.; Tümsek, F.; Chem. Eng. Res. Des. 2011, 89, 206. [Crossref]
  • 3
    Chang, C.-F.; Chang, C.-Y.; Tsai, W.-T.; J. Colloid Interface Sci. 2000, 232, 45. [Crossref]
  • 4
    Rodríguez-Reinoso, F.; Molina-Sabio, M.; Carbon N. Y. 1992, 30, 1111. [Crossref]
  • 5
    Yang, K.; Peng, J.; Xia, H.; Zhang, L.; Srinivasakannan, C.; Guo, S.; J. Taiwan Inst. Chem. Eng. 2010, 41, 367. [Crossref]
  • 6
    Demiral, H.; Demiral, I.; Fatma, T.; Karabacako, B.; Surf. Interface Anal. 2008, 40, 616. [Crossref]
  • 7
    Ubago-Pérez, R.; Carrasco-Marín, F.; Fairén-Jiménez, D.; Moreno-Castilla, C.; Microporous Mesoporous Mater. 2006, 92, 64. [Crossref]
  • 8
    Román, S.; González, J. F.; González-García, C. M.; Zamora, F.; Fuel Process. Technol. 2008, 89, 715. [Crossref]
  • 9
    González, J. F.; Román, S.; Encinar, J. M.; Martínez, G.; J. Anal. Appl. Pyrolysis 2009, 85, 134. [Crossref]
  • 10
    Ahmedna, M.; Marshall, W. E.; Rao, R. M.; Bioresour. Technol. 2000, 71, 113. [Crossref]
  • 11
    Lua, A. C.; Yang, T.; Guo, J.; J. Anal. Appl. Pyrolysis 2004, 72, 279. [Crossref]
  • 12
    Bouchelta, C.; Medjram, M. S.; Bertrand, O.; Bellat, J.-P.; J. Anal. Appl. Pyrolysis 2008, 82, 70. [Crossref]
  • 13
    Alau, K. K.; Gimba, C. E.; Kagbu, J. A.; Arch. Appl. Sci. Res. 2010, 2, 456. [Crossref]
  • 14
    Luo, L.; Wu, X.; Li, Z.; Zhou, Y.; Chen, T.; Fan, M.; Zhao, W.; R. Soc. Open Sci. 2019, 6, 190523. [Crossref]
  • 15
    Nagalakshmi, T. V.; Emmanuel, K. A.; Suresh Babu, C.; Chakrapani, C.; Divakar, P. P.; Int. Lett. Chem., Phys. Astron. 2015, 54, 189. [Crossref]
  • 16
    Olivares-Marín, M.; Fernández-González, C.; Macías-García, A.; Gómez-Serrano, V.; Appl. Surf. Sci. 2006, 252, 5980. [Crossref]
  • 17
    Laksaci, H.; Khelifi, A.; Trari, M.; Addoun, A.; J. Cleaner Prod. 2017, 147, 254. [Crossref]
  • 18
    Hanum, F.; Bani, O.; Izdiharo, A. M.; IOP Conf. Ser.: Mater. Sci. Eng. 2017, 180, 012149. [Crossref]
  • 19
    Álvarez-Torrellas, S.; Peres, J. A.; Gil-Álvarez, V.; Ovejero, G.; García, J.; Chem. Eng. J. 2017, 320, 319. [Crossref]
  • 20
    To, M.-H.; Hadi, P.; Hui, C. W.; Lin, C. S. K.; McKay, G.; J. Mol. Liq. 2017, 241, 386. [Crossref]
  • 21
    Berrios, M.; Martín, M.; Martín, A.; J. Ind. Eng. Chem. 2012, 18, 780. [Crossref]
  • 22
    Yakout, S. M.; Sharaf El-Deen, G.; Arabian J. Chem. 2016, 9, S1155. [Crossref]
  • 23
    Yahya, M. A.; Al-Qodah, Z.; Ngah, C. W. Z.; Renewable Sustainable Energy Rev. 2015, 46, 218. [Crossref]
  • 24
    Chile Oliva; Informe Anual Mercado Nacional de Aceite de Oliva; 2019. [Link] accessed in August 2022
  • 25
    Foods from Chile; La Industria del Aceite de Oliva en Chile. Informe de Sustentabilidad Chile Oliva; Chile, 2019. [Link] accessed in August 2022
  • 26
    Fundación para la Innovación Agraria; Con Residuos del Aceite de Oliva Desarrollan Biocombustibles Sólidos y Carbones Descontaminantes; Chile, 2020. [Link] accessed in August 2022
  • 27
    Zabaniotou, A.; Stavropoulos, G.; Skoulou, V.; Bioresour. Technol. 2008, 99, 320. [Crossref]
  • 28
    Azbar, N.; Bayram, A.; Filibeli, A.; Muezzinoglu, A.; Sengul, F.; Ozer, A.; Crit. Rev. Environ. Sci. Technol. 2004, 34, 209. [Crossref]
  • 29
    Martínez, M. L.; Torres, M. M.; Guzmán, C. A.; Maestri, D. M.; Ind. Crops Prod. 2006, 23, 23. [Crossref]
  • 30
    Ioannidou, O.; Zabaniotou, A.; Renewable Sustainable Energy Rev. 2007, 11, 1966. [Crossref]
  • 31
    Wong, S.; Ngadi, N.; Inuwa, I. M.; Hassan, O.; J. Cleaner Prod. 2018, 175, 361. [Crossref]
  • 32
    González, M. T.; Rodríguez-Reinoso, F.; García, A. N.; Marcilla, A.; Carbon N. Y. 1997, 35, 159. [Crossref]
  • 33
    Enaime, G.; Ennaciri, K.; Ounas, A.; Baçaoui, A.; Seffen, M.; Selmi, T.; Yaacoubi, A.; J. Mater. Environ. Sci. 2017, 8, 4125. [Crossref]
  • 34
    Moreno-Castilla, C.; Carrasco-Marín, F.; López-Ramón, M. V.; Alvarez-Merino, M. A.; Carbon N. Y. 2001, 39, 1415. [Crossref]
  • 35
    Rodríguez-Reinoso, F.; Molina-Sabio, M.; González, M. T.; Carbon N. Y. 1995, 33, 15. [Crossref]
  • 36
    Baçaoui, A.; Yaacoubi, A.; Dahbi, A.; Bennouna, C.; Phan Tan Luu, R.; Maldonado-Hodar, F. J.; Rivera-Utrilla, J.; Moreno-Castilla, C.; Carbon N. Y. 2001, 39, 425. [Crossref]
  • 37
    Linares-Solano, A.; Rodriguez-Reinoso, F.; Molina-Sabio, M.; Lopez-Gonzalez, J. D.; Adsorpt. Sci. Technol. 1984, 1, 223. [Crossref]
  • 38
    Elmouwahidi, A.; Bailón-García, E.; Pérez-Cadenas, A. F.; Maldonado-Hódar, F. J.; Carrasco-Marín, F.; Electrochim. Acta 2017, 229, 219. [Crossref]
  • 39
    Molina-Sabio, M.; Rodríguez-Reinoso, F.; Colloids Surf., A 2004, 241, 15. [Crossref]
  • 40
    Mochizuki, T.; Kubota, M.; Matsuda, H.; D’Elia Camacho, L. F.; Fuel Process. Technol. 2016, 144, 164. [Crossref]
  • 41
    Petit, C.; Karwacki, C.; Peterson, G.; Bandosz, T. J.; J. Phys. Chem. C 2007, 111, 12705. [Crossref]
  • 42
    Busca, G.; Pistarino, C.; J. Loss Prev. Process Ind. 2003, 16, 157. [Crossref]
  • 43
    Guo, J.; Xu, W. S.; Chen, Y. L.; Lua, A. C.; J. Colloid Interface Sci. 2005, 281, 285. [Crossref]
  • 44
    Haimour, N. M.; Emeish, S.; Waste Manage. 2006, 26, 651. [Crossref]
  • 45
    Solum, M. S.; Pugmire, R. J.; Jagtoyen, M.; Derbyshire, F.; Carbon N. Y. 1995, 33, 1247. [Crossref]
  • 46
    Jagtoyen, M.; Derbyshire, F.; Carbon N. Y. 1998, 36, 1085. [Crossref]
  • 47
    Yang, T.; Lua, A. C.; J. Colloid Interface Sci. 2003, 267, 408. [Crossref]
  • 48
    Zhang, T.; Walawender, W. P.; Fan, L. T.; Fan, M.; Daugaard, D.; Brown, R. C.; Chem. Eng. J. 2004, 105, 53. [Crossref]
  • 49
    Lillo-Ródenas, M. A.; Cazorla-Amorós, D.; Linares-Solano, A.; Carbon N. Y. 2003, 41, 267. [Crossref]
  • 50
    Marsh, H.; Yan, D. S.; O’Grady, T. M.; Wennerberg, A.; Carbon N. Y. 1984, 22, 603. [Crossref]
  • 51
    ASTM D1506: Standard Test Methods or Carbon Black - Ash Content, West Conshohocken, 2020.
  • 52
    ISO 562: Hard Coal and Coke-Determination of Volatile Matter, Switzerland, 2010.
  • 53
    Leofanti, G.; Padovan, M.; Tozzola, G.; Venturelli, B.; Catal. Today 1998, 41, 207. [Crossref]
  • 54
    Blanco, E.; Sepulveda, C.; Cruces, K.; García-Fierro, J. L.; Ghampson, I. T.; Escalona, N.; Catal. Today 2020, 356, 376. [Crossref]
  • 55
    Gil, A.; Grange, P.; Colloids Surf., A 1996, 113, 39. [Crossref]
  • 56
    ASTM-D4607-14: Standard Test Method for Determination of Iodine Number of Activated Carbon, West Conshohocken, 2014.
  • 57
    Boehm, H. P.; Carbon N. Y. 1994, 32, 759. [Crossref]
  • 58
    Benzekri, M. B.; Benderdouche, N.; Bestani, B.; Douara, N.; Duclaux, L.; J. Mater. Environ. Sci. 2018, 9, 272. [Crossref]
  • 59
    Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 345, 402. [Crossref]
  • 60
    Djelloul, C.; Hamdaoui, O.; Desalin. Water Treat. 2015, 56, 2966. [Crossref]
  • 61
    Hashem, A.; Al-Kheraije, K. A.; J. Environ. Prot. 2013, 4, 280. [Crossref]
  • 62
    Mediavilla, I.; Barro, R.; Borjabad, E.; Peña, D.; Fernández, M. J.; Renewable Energy 2020, 160, 374. [Crossref]
  • 63
    Bartocci, P.; D’Amico, M.; Moriconi, N.; Bidini, G.; Fantozzi, F.; Energy Procedia 2015, 82, 374. [Crossref]
  • 64
    Moubarik, A.; Grimi, N.; Food Res. Int. 2015, 73, 169. [Crossref]
  • 65
    Bedoui, A.; Souissi-Najar, S.; Ouederni, A.; Int. J. Innovation Appl. Stud. 2016, 18, 938. [Crossref]
  • 66
    Minkova, V.; Razvigorova, M.; Bjornbom, E.; Zanzi, R.; Budinova, T.; Petrov, N.; Fuel Process. Technol. 2001, 70, 53. [Crossref]
  • 67
    Gebreegziabher, T. B.; Wang, S.; Nam, H.; J. Environ. Chem. Eng. 2019, 7, 103234. [Crossref]
  • 68
    Hu, Z.; Srinivasan, M. P.; Microporous Mesoporous Mater. 2001, 43, 267. [Crossref]
  • 69
    Molina-Sabio, M.; González, M. T.; Rodriguez-Reinoso, F.; Sepúlveda-Escribano, A.; Carbon N. Y. 1996, 34, 505. [Crossref]
  • 70
    El-Hendawi, A.; Appl. Surf. Sci. 2009, 255, 3723. [Crossref]
  • 71
    Salame, I. I.; Bandosz, T. J.; Ind. Eng. Chem. Res. 2000, 39, 301. [Crossref]
  • 72
    Savova, D.; Apak, E.; Ekinci, E.; Yardim, F.; Petrov, N.; Budinova, T.; Razvigorova, M.; Minkova, V.; Biomass Bioenergy 2001, 21, 133. [Crossref]
  • 73
    Sun, K.; Jiang, J. C.; Biomass Bioenergy 2010, 34, 539. [Crossref]
  • 74
    Budinova, T.; Ekinci, E.; Yardim, F.; Grimm, A.; Björnbom, E.; Minkova, V.; Goranova, M.; Fuel Process. Technol. 2006, 87, 899. [Crossref]
  • 75
    László, K.; Microporous Mesoporous Mater. 2005, 80, 205. [Crossref]
  • 76
    Radovic, L.; Surface Chemistry of Activated Carbon Materials: State of the Art and Implications for Adsorption, Surfaces of Nanoparticles and Porous Materials; Schwarz, J.; Contescu, C., eds.; Marcel Dekker INC.: New York, 1999.
  • 77
    Guo, J.; Lua, A. C.; Mater. Lett. 2002, 55, 334. [Crossref]
  • 78
    Termoul, M.; Bestani, B.; Benderdouche, N.; Belhakem, M.; Adsorpt. Sci. Technol. 2006, 24, 375. [Crossref]
  • 79
    Esteves, B. M.; Morales-Torres, S.; Maldonado-Hódar, F. J.; Madeira, L. M.; Chem. Eng. J. 2021, 411, 128451. [Crossref]
  • 80
    Youssef, A.; El-Nabarawy, T.; El-Shafey, E.; Carbon Lett. 2006, 7, 1. [Crossref]
  • 81
    Şirazi, M.; Aslan, S.; Chem. Eng. Commun. 2021, 208, 1479. [Crossref]
  • 82
    Choi, S.; Choi, W.; Kim, S.; Lee, S.-Y.; Noh, I.; Kim, C. W.; Biomater. Res. 2014, 18, 6. [Crossref]
  • 83
    Dastgheib, S. A.; Rockstraw, D. A.; Carbon N. Y. 2001, 39, 1849. [Crossref]
  • 84
    Toles, C. A.; Marshall, W. E.; Johns, M. M.; Carbon N. Y. 1999, 37, 1207. [Crossref]
  • 85
    Mahmood, T.; Ali, R.; Naeem, A.; Hamayun, M.; Aslam, M.; Process Saf. Environ. Prot. 2017, 109, 548. [Crossref]
  • 86
    Benaddi, H.; Bandosz, T. J.; Jagiello, J.; Schwarz, J. A.; Rouzaud, J. N.; Carbon N. Y. 2000, 38, 669. [Crossref]
  • 87
    Gonçalves, M.; Sánchez-García, L.; de Oliveira Jardim, E.; Silvestre-Albero, J.; Rodríguez-Reinoso, F.; Environ. Sci. Technol. 2011, 45, 10605. [Crossref]
  • 88
    Figueiredo, J. L.; Pereira, M. F. R.; Freitas, M. M. A.; Órfão, J. J. M.; Carbon N. Y. 1999, 37, 1379. [Crossref]
  • 89
    Figueiredo, J. L.; Pereira, M. F. R.; Catal. Today 2010, 150, 2. [Crossref]
  • 90
    Ro, K. S.; Lima, I. M.; Reddy, G. B.; Jackson, M. A.; Gao, B.; Agriculture 2015, 5, 991. [Crossref]
  • 91
    Zhou, J. H.; Sui, Z.-J.; Zhu, J.; Li, P.; Chen, D.; Dai, Y.-C.; Yuan, W.-K.; Carbon N. Y. 2007, 45, 785. [Crossref]
  • 92
    Domingo-García, M.; Groszek, A. J.; López-Garzón, F. J.; Pérez-Mendoza, M.; Appl. Catal., A 2002, 233, 141. [Crossref]
  • 93
    Wu, X.; Radovic, L. R.; Carbon N. Y. 2006, 44, 141. [Crossref]
  • 94
    Song, J.; Zou, W.; Bian, Y.; Su, F.; Han, R.; Desalination 2011, 265, 119. [Crossref]
  • 95
    Cuhadaroglu, D.; Uygun, O. A.; Afr. J. Biotechnol. 2008, 7, 3706. [Crossref]
  • 96
    Han, R.; Zhang, L.; Song, C.; Zhang, M.; Zhu, H.; Zhang, L. J.; Carbohydr. Polym. 2010, 79, 1140. [Crossref]
  • 97
    Fanning, P. E.; Vannice, M. A.; Carbon N. Y. 1993, 31, 721. [Crossref]
  • 98
    Guo, J.; Lua, A. C.; Microporous Mesoporous Mater. 1999, 32, 111. [Crossref]
  • 99
    Pérez Marín, A. B.; Aguilar, M. I.; Meseguer, V. F.; Ortuño, J. F.; Sáez, J.; Lloréns, M.; Chem. Eng. J. 2009, 155, 199. [Crossref]
  • 100
    Asada, T.; Ohkubo, T.; Kawata, K.; Oikawa, K.; J. Heal. Sci. 2006, 52, 585. [Crossref]
  • 101
    Helminen, J.; Helenius, J.; Paatero, E.; Turunen, I.; J. Chem. Eng. Data 2001, 46, 391. [Crossref]
  • 102
    Yeom, C.; Kim, Y.; Environ. Eng. Res. 2017, 22, 401. [Crossref]
  • 103
    Rieth, A. J.; Dinca, M.; J. Am. Chem. Soc. 2018, 140, 3461. [Crossref]
  • 104
    Park, S.-J.; Kim, K.-D.; J. Colloid Interface Sci. 1999, 212, 186. [Crossref]
  • 105
    Shan, X.-m.; Zhu, S.-q.; Zhang, W.-h.; J. China Univ. Min. Technol. 2008, 18, 261. [Crossref]
  • 106
    Saha, D.; Deng, S.; J. Colloid Interface Sci. 2010, 348, 615. [Crossref]
  • 107
    Sangon, S.; Hunt, A. J.; Attard, T. M.; Mengchang, P.; Ngernyen, Y.; Supanchaiyamat, N.; J. Cleaner Prod. 2018, 172, 1128. [Crossref]
  • 108
    Liang, P.; Yu, H.; Huang, J.; Zhang, Y.; Cao, H.; MATEC Web Conf. 2016, 67, article 07006. [Crossref]
  • 109
    Hayhurst, D. T.; Chem. Eng. Commun. 1980, 4, 729. [Crossref]

Edited by

Editor handled this article: Jaísa Fernandes Soares

Publication Dates

  • Publication in this collection
    10 Feb 2023
  • Date of issue
    2023

History

  • Received
    27 Mar 2022
  • Accepted
    10 Aug 2022
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br