Acessibilidade / Reportar erro

Microporous Nitrogen-Doped Activated Biochars Derived from Corn: Use of Husk Waste and Urea for CO2 Capture

Abstract

The increase in anthropogenic activity over time has led to an exponential increase in greenhouse gas emissions, especially CO2. Proftable technologies for CO2 capture and separation are conspicuous, and porous biochars derived from biomass waste can be a useful solution. Herein, we produced activated nitrogen-doped biochars for CO2 capture from corn husk waste, urea and K2CO3, named N-Bio-X (X = 600, 700, and 800 °C). N-Bio-X exhibited microporosity and different nitrogen contents and thus played an important role in the adsorption of CO2. N-Bio-700 exhibited the highest CO2 adsorption capacity, fastest adsorption kinetics and excellent stability after multiple adsorption-desorption cycles. N-Bio-600 showed excellent CO2/N2 selectivity, induced by nitrogen sites, particularly pyridinic and graphitic nitrogen. The cost-effectiveness of the raw material, coupled with its high adsorption capacity, rapid kinetics, and stable properties, provided highly promising N-doped biochars for practical implementation in CO2 capture and separations in postcombustion processes.

Keywords:
biomass; corn husk; urea; nitrogen-doped biochar; CO2 capture and separation; postcombustion


Introduction

According to the Intergovernmental Panel on Climate Change (IPCC),11 Pachauri, R. K.; Allen, M. R.; Barros, V. R.; Broome, J.; Cramer, W.; Christ, R.; Church, J. A.; Clarke, L.; Dahe, Q.; Dasgupta, P.; Dubash, N. K.; Edenhofer, O.; Elgizouli, I.; Field, C. B.; Forster, P.; Friedlingstein, P.; Fuglestvedt, J.; Gomez-Echeverri, L.; Hallegatte, S.; Hegerl, G.; Howden, M.; Jiang, K.; Cisneros, B. J.; Kattsov, V.; Lee, H.; Mach, K. J.; Marotzke, J.; Mastrandrea, M. D.; Meyer, L.; Minx, J.; Mulugetta, Y.; O’Brien, K.; Oppenheimer, M.; Pereira, J. J.; Pichs-Madruga, R.; Plattner, G.-K.; Pörtner, H.-O.; Power, S. B.; Preston, B.; Ravindranath, N. H.; Reisinger, A.; Riahi, K.; Rusticucci, M.; Scholes, R.; Seyboth, K.; Sokona, Y.; Stavins, R.; Stocker, T. F.; Tschakert, P.; van Vuuren, D.; van Ypersele, J.-P.; Climate Change 2014: Mitigation of Climate Change, IPCC Fifth Assessment Report, Cambridge University Press, 2015. [Crossref]
Crossref...
the burning of fossil fuels is responsible for approximately 70% of global CO2 emissions. The environmental consequences of these emissions include rising global temperatures, ocean acidifcation and rising sea levels, all of which affect biodiversity and food security. To mitigate these impacts, the use of CO2 capture technologies in industry and even in treating the atmosphere has been identifed as a promising solution.11 Pachauri, R. K.; Allen, M. R.; Barros, V. R.; Broome, J.; Cramer, W.; Christ, R.; Church, J. A.; Clarke, L.; Dahe, Q.; Dasgupta, P.; Dubash, N. K.; Edenhofer, O.; Elgizouli, I.; Field, C. B.; Forster, P.; Friedlingstein, P.; Fuglestvedt, J.; Gomez-Echeverri, L.; Hallegatte, S.; Hegerl, G.; Howden, M.; Jiang, K.; Cisneros, B. J.; Kattsov, V.; Lee, H.; Mach, K. J.; Marotzke, J.; Mastrandrea, M. D.; Meyer, L.; Minx, J.; Mulugetta, Y.; O’Brien, K.; Oppenheimer, M.; Pereira, J. J.; Pichs-Madruga, R.; Plattner, G.-K.; Pörtner, H.-O.; Power, S. B.; Preston, B.; Ravindranath, N. H.; Reisinger, A.; Riahi, K.; Rusticucci, M.; Scholes, R.; Seyboth, K.; Sokona, Y.; Stavins, R.; Stocker, T. F.; Tschakert, P.; van Vuuren, D.; van Ypersele, J.-P.; Climate Change 2014: Mitigation of Climate Change, IPCC Fifth Assessment Report, Cambridge University Press, 2015. [Crossref]
Crossref...
The European Union has set a CO2 capture target of 43 million tons by 2030. Therefore, a signifcant increase in CO2 capture capacity is needed to reach these targets.22 Global CCS Institute, Global Status of CCS Report: 2020, https://www.globalccsinstitute.com/resources/publications-reports-research/global-status-of-ccs-report-2020/, accessed in February 2024.
https://www.globalccsinstitute.com/resou...
,33 Chagas, J. A. O.; Marciniak, A. A.; Mota, C. J. A.; J. Braz. Chem. Soc. 2022, 33, 801. [Crossref]
Crossref...

CO2 capture and separation can be performed by chemical absorption using amine aqueous solutions,44 dos Santos, T. C.; Lage, M. R.; da Silva, A. F. M.; Fernandes, T. S.; Carneiro, J. W. M.; Ronconi, C. M.; J. CO2 Util. 2022, 61, 102054. [Crossref]
Crossref...
,55 Furtado, I. O.; dos Santos, T. C.; Vasconcelos, L. F.; Costa, L. T.; Fiorot, R. G.; Ronconi, C. M.; Carneiro, J. W. M.; Chem. Eng. J. 2021, 408, 128002. [Crossref]
Crossref...
carbonate solutions,66 de Vasconcelos, S. F.; Carneiro, L. O.; Brito, R. P.; Brito, K. D.; J. Braz. Chem. Soc. 2023, 34, 441. [Crossref]
Crossref...
ionic liquids,77 Aquino, A. S.; Bernard, F. L.; Vieira, M. O.; Borges, J. V.; Rojas, M. F.; Vecchia, F. D.; Ligabue, R. A.; Seferin, M.; Menezes, S.; Einloft, S.; J. Braz. Chem. Soc. 2014, 25, 2251. [Crossref]
Crossref...
adsorption on porous solid materials,88 Freitas, S. K. S.; Oliveira, F. L.; dos Santos, T. C.; Hisse, D.; Merlini, C.; Ronconi, C. M.; Esteves, P. M.; Chem. Eur. J. 2021, 27, 2342. [Crossref]
Crossref...
,99 dos Santos, T. C.; Ronconi, C. M.; J. CO2 Util. 2017, 20, 292. [Crossref]
Crossref...
membrane separations1010 Wang, X.; Song, C.; Front. Energy Res. 2020, 8, 265. [Crossref]
Crossref...
and cryogenic distillation.1111 Quang, D. V.; Milani, D.; Abu Zahra, M.; Int. J. Greenhouse Gas Control 2023, 124, 103862. [Crossref]
Crossref...
Amine aqueous solutions for CO2 absorption have been applied in coal-fred power generation and natural gas purifcation.1212 Ribeiro, M. G.; Hisse, D.; Prado, M. L.; dos Santos, T. C.; Ronconi, C. M.; Rev. Virtual Quim. 2022, 14, 517. [Crossref]
Crossref...
Although this method is very effcient, its high operating and maintenance costs, as well as the potential environmental impacts of its toxic byproducts arising from amine degradation, could restrict its extensive use for CO2 capture in other applications.44 dos Santos, T. C.; Lage, M. R.; da Silva, A. F. M.; Fernandes, T. S.; Carneiro, J. W. M.; Ronconi, C. M.; J. CO2 Util. 2022, 61, 102054. [Crossref]
Crossref...
Therefore, signifcant effort has been made to develop environmentally friendly methods for low-cost CO2 capture. CO2 adsorption has several advantages over other carbon capture technologies because it can be used in a variety of settings, such as power plants, industrial processes, and transportation, and it is easily integrated into existing structures. Moreover, these materials can be combined with other carbon capture technologies to improve their effciency.

There are a variety of adsorbent materials for CO2 capture, including carbon-based materials,99 dos Santos, T. C.; Ronconi, C. M.; J. CO2 Util. 2017, 20, 292. [Crossref]
Crossref...
,1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
ordered porous silicas,1414 dos Santos, T. C.; Bourrelly, S.; Llewellyn, P. L.; Carneiro, J. W. D. M.; Ronconi, C. M.; Phys. Chem. Chem. Phys. 2015, 17, 11095. [Crossref]
Crossref...
1515 Mello, M. R.; Phanon, D.; Silveira, G. Q.; Llewellyn, P. L.; Ronconi, C. M.; Microporous Mesoporous Mater. 2011, 143, 174. [Crossref]
Crossref...
1616 dos Santos, T. C.; Ronconi, C. M.; Rev. Virtual Quim. 2014, 6, 112. [Crossref]
Crossref...
zeolites,1717 Babu, P. E.; Tejes, P. K. S.; Naik, B. K.; Carbon Capture Sci. Technol. 2023, 7, 100103. [Crossref]
Crossref...
,1818 Najafi, A. M.; Soltanali, S.; Khorashe, F.; Ghassabzadeh, H.; Chemosphere 2023, 324, 138275. [Crossref]
Crossref...
metal-organic frameworks,1919 Dinda, S.; Mater. Today Commun. 2023, 35, 105927. [Crossref]
Crossref...
and covalent-organic frameworks.88 Freitas, S. K. S.; Oliveira, F. L.; dos Santos, T. C.; Hisse, D.; Merlini, C.; Ronconi, C. M.; Esteves, P. M.; Chem. Eur. J. 2021, 27, 2342. [Crossref]
Crossref...
These materials usually have high surface areas and pore volumes and tunable pore sizes. The selectivities and adsorption capacities of ordered porous silicas,1414 dos Santos, T. C.; Bourrelly, S.; Llewellyn, P. L.; Carneiro, J. W. D. M.; Ronconi, C. M.; Phys. Chem. Chem. Phys. 2015, 17, 11095. [Crossref]
Crossref...
1515 Mello, M. R.; Phanon, D.; Silveira, G. Q.; Llewellyn, P. L.; Ronconi, C. M.; Microporous Mesoporous Mater. 2011, 143, 174. [Crossref]
Crossref...
1616 dos Santos, T. C.; Ronconi, C. M.; Rev. Virtual Quim. 2014, 6, 112. [Crossref]
Crossref...
zeolites,2020 Salih, H. A.; Alkhatib, I. I. I.; Zahra, M. A.; Vega, L. F.; J. CO2 Util. 2023, 68, 102383. [Crossref]
Crossref...
,2121 Short, G. N.; Burentugs, E.; Proaño, L.; Moon, H. J.; Rim, G.; Nezam, I.; Korde, A.; Nair, S.; Jones, C. W.; JACS Au 2023, 3, 62. [Crossref]
Crossref...
metal-organic frameworks,1919 Dinda, S.; Mater. Today Commun. 2023, 35, 105927. [Crossref]
Crossref...
,2222 Huang, Z.; Ying, L.; Gong, F.; Lu, J.; Wang, W.; Ding, J.; Yan, J.; J. Environ. Chem. Eng. 2023, 11, 109739. [Crossref]
Crossref...
,2323 Ahn, H.; Xu, R.; Kim, S.; Lee, J. W.; Kang, Y. T.; J. Environ. Chem. Eng. 2023, 11, 109723. [Crossref]
Crossref...
and covalent-organic frameworks toward CO2 can be improved by surface or pore chemical functionalization with amino groups.

Carbon-based materials, particularly biochars, are promising adsorbents for CO2 capture because they can be obtained from biomass waste, which reduces the costs of production and is easily scalable.2424 Zhang, X.; Cao, L.; Xiang, W.; Xu, Y.; Gao, B.; Sep. Purif. Technol. 2022, 295, 121295. [Crossref]
Crossref...
2525 Bamdad, H.; Hawboldt, K.; Macquarrie, S.; Energy Fuels 2018, 32, 11742. [Crossref]
Crossref...
2626 Chatterjee, R.; Sajjadi, B.; Chen, W. Y.; Mattern, D. L.; Egiebor, N. O.; Hammer, N.; Raman, V.; Energy Fuels 2019, 33, 2366. [Crossref]
Crossref...
2727 Wu, B.; Song, X.; Zheng, D.; Tan, Q.; Yao, Y.; Liu, F. Q.; ACS Appl. Mater. Interfaces 2023, 15, 20325. [Crossref]
Crossref...
2828 Li, D.; Chen, W.; Wu, J.; Jia, C. Q.; Jiang, X.; J. Mater. Chem. A 2020, 8, 24977. [Crossref]
Crossref...
Currently, agricultural residues, including eucalyptus bark,2929 Kortunov, P. V.; Siskin, M.; Paccagnini, M.; Thomann, H.; Energy Fuels 2016, 30, 1223. [Crossref]
Crossref...
banana and grapefruit peel,3030 Li, W.; Tu, W.; Cheng, J.; Yang, F.; Wang, X.; Li, L.; Shang, D.; Zhou, X.; Yu, C.; Yuan, A.; Pan, J.; Sep. Purif. Technol. 2022, 282, 120001. [Crossref]
Crossref...
rice husk3131 Li, D.; Ma, T.; Zhang, R.; Tian, Y.; Qiao, Y.; Fuel 2015, 139, 68. [Crossref]
Crossref...
and coconut,3232 Yue, L.; Xia, Q.; Wang, L.; Wang, L.; DaCosta, H.; Yang, J.; Hu, X.; J. Colloid Interface Sci. 2018, 511, 259. [Crossref]
Crossref...
,3333 Liang, W.; Luo, Z.; Zhou, D.; Liu, Z.; Wei, X.; Cai, W.; Chem. Eng. J. 2023, 461, 141980. [Crossref]
Crossref...
peanut,3434 Luo, Z.; Peng, X.; Liang, W.; Zhou, D.; Dang, C.; Cai, W.; Bioresour. Technol. 2023, 388, 129762. [Crossref]
Crossref...
,3535 Peng, X.; Luo, Z.; Xie, H.; Liang, W.; Luo, J.; Dang, C.; Wang, A.; Hu, L.; Yu, X.; Cai, W.; Appl. Surf. Sci. 2022, 606, 154859. [Crossref]
Crossref...
and macadamia nutshells,3636 Bai, J.; Huang, J.; Yu, Q.; Demir, M.; Kilic, M.; Altay, B. N.; Hu, X.; Wang, L.; Fuel Process Technol. 2023, 249, 107854. [Crossref]
Crossref...
have been used as feedstocks to produce porous carbon-based materials for the adsorption of organic pollutants, heavy metals and CO2.3737 Khan, M. H.; Akash, N. M.; Akter, S.; Rukh, M.; Nzediegwu, C.; Islam, M. S.; J. Environ. Manage. 2023, 338, 117825. [Crossref]
Crossref...
,3838 Hanif, A.; Aziz, M. A.; Helal, A.; Abdelnaby, M. M.; Khan, A.; Theravalappil, R.; Khan, M. Y.; ACS Omega 2023, 8, 36228. [Crossref]
Crossref...
Biochars are thermically stable and highly microporous, have high CO2 adsorption capacities at low pressures, exhibit fast CO2 adsorption/desorption kinetics and are therefore good candidates for CO2 capture and separation in postcombustion processes.3939 Ghanbarpour Mamaghani, Z.; Hawboldt, K. A.; MacQuarrie, S.; J. Environ. Chem. Eng. 2023, 11, 109643. [Crossref]
Crossref...
,4040 Zaker, A.; ben Hammouda, S.; Sun, J.; Wang, X.; Li, X.; Chen, Z.; J. Environ. Chem. Eng. 2023, 11, 109741. [Crossref]
Crossref...
Additionally, the affnities and selectivities of these materials for CO2 can be improved by the introduction of nitrogen atoms into their frameworks through copyrolysis with nitrogen-containing compounds such as melamine,4141 Adio, S. O.; Ganiyu, S. A.; Usman, M.; Abdulazeez, I.; Alhooshani, K.; Chem. Eng. J. 2020, 382, 122964. [Crossref]
Crossref...
ethylenediamine,4242 Ma, X.; Su, C.; Liu, B.; Wu, Q.; Zhou, K.; Zeng, Z.; Li, L.; Sep. Purif. Technol. 2021, 259, 118065. [Crossref]
Crossref...
sodium amide,4343 Rao, L.; Liu, S.; Wang, L.; Ma, C.; Wu, J.; An, L.; Hu, X.; Chem. Eng. J. 2019, 359, 428. [Crossref]
Crossref...
and urea.4444 Luo, J.; Liu, B.; Shi, R.; Guo, Y.; Feng, Q.; Liu, Z.; Li, L.; Norinaga, K.; Microporous Mesoporous Mater. 2021, 327, 111404. [Crossref]
Crossref...
The nitrogen species are more basic than the carbon atoms and they improve the CO2 adsorption capacities at low pressures and thus increase the CO2 selectivity of the gas mixture. The adsorption capacity can also be improved by the addition of KOH to the biomass during pyrolysis to increase the surface area and pore volume of the resulting material.4545 Chun, S.-E.; Whitacre, J. F.; Microporous Mesoporous Mater. 2017, 251, 34. [Crossref]
Crossref...
At temperatures higher than 600 °C, KOH reacts with the carbon framework to remove carbon and create pores.4040 Zaker, A.; ben Hammouda, S.; Sun, J.; Wang, X.; Li, X.; Chen, Z.; J. Environ. Chem. Eng. 2023, 11, 109741. [Crossref]
Crossref...
However, alkaline KOH is a corrosive and hazardous chemical, and other activating reagents, such as K2CO3, have been proposed to create porosity.4646 Hayashi, J.; Yamamoto, N.; Horikawa, T.; Muroyama, K.; Gomes, V. G.; J. Colloid Interface Sci. 2005, 281, 437. [Crossref]
Crossref...

In this work, N-doped activated biochars for CO2 capture (N-Bio-X, X = 600, 700 and 800 °C) were prepared from corn husk as a feedstock, urea as a nitrogen source and K2CO3 as an activating agent. As the second largest agricultural area in the world, corn produces a large amount of corn husk waste and can be used as a biomass source to produce biochars.4747 Ratna, A. S.; Ghosh, A.; Mukhopadhyay, S.; J. Cleaner Prod. 2022, 371, 133563. [Crossref]
Crossref...
This practice can promote waste reduction by developing sustainable and effcient materials, thereby safeguarding our natural resources.4848 Fattahi, M.; Taban, E.; Soltani, P.; Berardi, U.; Khavanin, A.; Zaroushani, V.; J. Building Engineering 2023, 77, 107468. [Crossref]
Crossref...
However, to our knowledge, few studies have explored this biomass as a feedstock for biochar preparation for use as an adsorbent for CO2 capture. Because urea is an inexpensive chemical, it has become an interesting nitrogen-containing compound with which to dope biochar from corn husks. Therefore, we investigated the effects of activation temperature on the surface area, pore volume, pore size and nitrogen percentage of N-Bio-X and correlated these properties with the CO2 adsorption capacity, thermodynamics, and selectivity. Kinetic and cyclic adsorption-desorption studies were also carried out to understand the behavior of the synthesized biochars under CO2 flow.

Experimental

Materials

The following reactants and solvents were used: potassium carbonate (K2CO3, P.A., Sigma-Aldrich, St. Louis, USA), urea (NH2)2CO, 99.5%, Sigma-Aldrich, St. Louis, USA), ethanol (EtOH, 95%, Isofar, Duque de Caxias, Brazil) and hydrochloric acid (HCl, 37%, Isofar, Duque de Caxias, Brazil). Corn husks were purchased from a local market.

Preparation of biochar from corn husk biomass (Bio-500)

The corn husks were washed with water, dried at 100 °C for 2 h and ground in a conventional blender until a homogeneous powder was formed. Then, the powder (535 g) was heated at 500 °C (10 °C min-1) under a static air atmosphere for 2 h. The obtained biochar was named Bio-500 (10 g) and used as a precursor to obtain nitrogen-doped biochars. Elemental analysis (CHN) of the Bio-500 samples revealed the following: C, 73.20%; H, 3.16%; and N, 1.71%.

Preparation of N-doped and activated carbons (N-Bio-X, X = 600, 700 and 800 °C)

The preparation of the biochars is schematically represented in Figure 1.

Figure 1.
Representation of the procedure employed for the preparation of the biochars.

In a porcelain mortar, 1.0 g of the Bio-500 precursor, 1.0 g of urea and 1.5 g of K2CO3 with a mass ratio of 1:1:1.5 were ground to obtain a homogeneous mixture, which was subsequently added to a 100 mL beaker containing 80 mL of EtOH. The mixture was heated to 80 °C and stirred magnetically until solvent evaporation. The black solid was dried at 90 °C in a drying oven. The resulting solid was introduced in a vertical furnace on a glass wool surface and heated at 600, 700 or 800 °C under a N2 atmosphere (150 mL min1) at a heating rate of 5 °C min1 for 2 h. The resulting products were washed with 1 mol L1 HCl to remove the remaining unreacted K2CO3 and then with distilled water up to neutral pH. The solids were dried at 110 °C overnight to form nitrogen-doped biochars, which were named N-Bio-X (X = 600, 700 and 800 °C), yielding 0.33 g (N-Bio-600), 0.27 g (N-Bio-700) and 0.20 g (N-Bio-800) of each material. Elemental analyses (CHNs) were performed for N-Bio-600: C, 62.01%; H, 2.91%; N, 3.58%; N-Bio-700: C, 57.12%; H, 3.43%; N, 2.12%; N-Bio-800: C, 56.74%; H, 4.05%; and N, 0.74%.

Characterization techniques

The CHN contents were determined with a PerkinElmer (Shelton, USA) CHN 240 C Elemental Analyzer (IQ-USP). Powder X-ray diffraction (PDRX) patterns were acquired on a Minifex Rigaku II (Tokyo, Japan) diffractometer (Cu Kα radiation = 1.5418 Å) with a 2θ range of 10-60°. Raman spectra were collected on a Witec Alpha 300 spectrometer (Ulm, Germany) using an excitation laser at λ= 532 nm and a 50× lens glass. X-ray photoelectron spectroscopy (XPS) was performed with a Thermo Scientifc Escalab 250Xi (Waltham, USA) system with monochromated Al Kα radiation (1486.6 eV). The survey spectra were recorded using 0.2 eV steps and a 100 eV analyzer pass energy. The high-resolution N 1s binding energies were measured with a 25 eV analyzer pass energy with 0.2 eV steps. Deconvolutions of the signals centered at 400.1 eV (N 1s) were carried out with CasaXPS software.4949 Ayiania, M.; Smith, M.; Hensley, A. J. R.; Scudiero, L.; McEwen, J. S.; Garcia-Perez, M.; Carbon 2020, 162, 528. [Crossref]
Crossref...
Nitrogen adsorption/desorption isotherms were measured on a Micromeritics ASAP 2020 (Norcross, USA) system at -196 °C. For these measurements, the samples were pretreated at 180 °C under vacuum for 24 h. The Brunauer-Emmett-Teller (BET) equation and the Horvath-Kawazoe (HK) method were used to determine the specifc surface areas and cumulative pore volumes, respectively. Transmission electron microscopy (TEM) images were obtained with a probe-corrected FEI Titan 80-300 (Hillsboro, USA) transmission electron microscope (80 kV). The images were acquired in scanning TEM (STEM) mode using a high-angle annular dark-field (HAADF) detector. The samples were prepared by dispersing 1.0 mg of the biochars in isopropyl alcohol (3 mL) in an ultrasonic bath for 10 min. Then, one drop of the dispersion was placed on a lacey carbon copper grid (300 mesh) and dried at room temperature.

CO2 adsorption experiments

CO2 isotherms were measured in the low-pressure regime (0-1 bar) at three different temperatures, 25, 50 and 70 °C, with a Micromeritics ASAP 2020 (Norcross, USA) system. The equilibrium parameters were obtained using three adsorption mathematical models, the Langmuir, Freundlich and Toth equations, with the dataft function of the free software Scilab.5050 Scilab, 6.10; French Institute for Research in Computer Science and Control, Vélizy-Villacoublay, France, 1984. The best model was selected based on the normalized standard deviations. The adsorption enthalpy was calculated from the Clausius-Clapeyron equation. The same temperature and pressure conditions used to obtain CO2 isotherms were used to obtain N2 isotherms and determine the CO2/N2 selectivities (IASTCO2/N2). The selectivity IASTCO2/N2 was calculated using the ideal adsorption solution theory (IAST) derived from Henry’s law and the Gibbs isotherm equations.5151 Cessford, N. F.; Seaton, N. A.; Düren, T.; Ind. Eng. Chem. Res. 2012, 51, 4911. [Crossref]
Crossref...

Kinetic CO2 uptake adsorption experiments

The kinetic isotherms for CO2 uptake were obtained on a Shimadzu DTG-60/60-H TG (Japan) apparatus based on the procedure reported by Yang et al.5252 Yang, J.; Yue, L.; Hu, X.; Wang, L.; Zhao, Y.; Lin, Y.; Sun, Y.; DaCosta, H.; Guo, L.; Energy Fuels 2017, 31, 4287. [Crossref]
Crossref...
The samples (7.0 mg) were placed in alumina crucibles and pretreated at 120 °C under a N2 fow of 50 mL min-1 for 1 h or until the mass stabilized. The temperature was cooled to 30 °C, and CO2 was added to the sample (50 mL min-1), whereas a N2 fow (50 mL min-1) was passed through the balance. The CO2 uptake adsorption was measured until no mass variation was detected (equilibrium), and then the CO2 fow was switched off to allow CO2 desorption until the initial mass was reached. The desorption time was held the same for each cycle. This procedure was repeated seven times.

Results and Discussion

Syntheses of the nitrogen-doped biochars: Bio-500 and N-Bio-X (X = 600, 700 and 800 °C)

Biochar Bio-500 was obtained by heating ground corn husk waste at 500 °C under a static air atmosphere for 2 h. The percentage of nitrogen in Bio-500, determined by elemental analysis, was 1.71%. Corn husks usually contain proteins and lignin-containing nitrogen compounds, which can be incorporated into biochar as nitrogen atoms during thermal decomposition.5353 Yang, X.; Han, F.; Xu, C.; Jiang, S.; Huang, L.; Liu, L.; Xia, Z.; Ind. Crops Prod. 2017, 109, 241. [Crossref]
Crossref...
5454 Del Río, J. C.; Rencoret, J.; Gutiérrez, A.; Kim, H.; Ralph, J.; J. Agric. Food Chem. 2018, 66, 4402. [Crossref]
Crossref...
5555 Yeasmin, M. S.; Mondal, M. I. H.; Int. J. Biol. Macromol. 2015, 80, 725. [Crossref]
Crossref...
We added more nitrogen atoms to Bio-500 by using urea as the nitrogen source and K2CO3 to activate the material at different calcination temperatures, resulting in nitrogen-doped biochars N-Bio-X (X = 600, 700 and 800 °C).

At high temperatures, K2CO3 decomposed into CO2 and K2O, etched the carbon atoms in the lattice of Bio-500 and produced microporous structures based on the reactions shown in equations 1, 2, 3. At temperatures higher than 700 °C and in the presence of K2CO3, the pore volume and/or size increase.4040 Zaker, A.; ben Hammouda, S.; Sun, J.; Wang, X.; Li, X.; Chen, Z.; J. Environ. Chem. Eng. 2023, 11, 109741. [Crossref]
Crossref...
,5656 Singh, G.; Kim, I. Y.; Lakhi, K. S.; Srivastava, P.; Naidu, R.; Vinu, A.; Carbon 2017, 116, 448. [Crossref]
Crossref...

(1) K 2 CO 3 ( s ) K 2 O ( s ) + CO 2 ( g )

(2) C ( s ) + CO 2 ( g ) 2 CO ( g )

(3) K 2 O ( s ) +C ( s ) 2 K ( s ) + CO ( g )

The N-doping process occurs via thermal decomposition of urea and is a complex process depending on the heating temperature.5757 Zhu, N.; Qian, F.; Xu, X.; Wang, M.; Teng, Q.; Materials 2021, 14, 6190. [Crossref]
Crossref...
,5858 Tischer, S.; Börnhorst, M.; Amsler, J.; Schoch, G.; Deutschmann, O.; Phys. Chem. Chem. Phys. 2019, 21, 16785. [Crossref]
Crossref...
Briefy, at temperatures above 133 °C, urea decomposes into ammonia, biuret, triuret, and isocyanic acid.5858 Tischer, S.; Börnhorst, M.; Amsler, J.; Schoch, G.; Deutschmann, O.; Phys. Chem. Chem. Phys. 2019, 21, 16785. [Crossref]
Crossref...
The latter compound can further decompose into cyanuric acid, which sublimes at temperatures above 300 °C. The percentages of nitrogen atoms incorporated into the biochars, as determined by elemental analyses, were 3.58, 2.12, and 0.74% for the N-Bio-600, N-Bio-700, and N-Bio-800 samples, respectively. The percentage of nitrogen atoms decreased as the activation temperature increased, which may be attributed to thermal reactions of the incorporated nitrogen atoms.

Wang and co-workers 3636 Bai, J.; Huang, J.; Yu, Q.; Demir, M.; Kilic, M.; Altay, B. N.; Hu, X.; Wang, L.; Fuel Process Technol. 2023, 249, 107854. [Crossref]
Crossref...
,5959 Bai, J.; Huang, J.; Jiang, Q.; Jiang, W.; Demir, M.; Kilic, M.; Altay, B. N.; Wang, L.; Hu, X.; Colloids Surf., A 2023, 674, 131916. [Crossref]
Crossref...
,6060 Bai, J.; Huang, J.; Yu, Q.; Demi, M.; Akgul, E.; Altay, B. N.; Hu, X.; Wang, L.; Front Chem. Sci. Eng. 2023, 17, 1122. [Crossref]
Crossref...
,6161 Bai, J.; Shao, J.; Yu, Q.; Demir, M.; Altay, B. N.; Ali, T. M.; Jiang, Y.; F, L. W.; Hu, X.; Chem. Eng. J. 2024, 479, 147667. [Crossref]
Crossref...
reported the effect of the activator amount on the development of the material porosity. Increasing the activator ratio in the feedstock resulted in higher values of specifc surface area and pore volume. Nonetheless, Tanaka and co-workers6262 Khuong, D. A.; Trinh, K. T.; Nakaoka, Y.; Tsubota, T.; Tashima, D.; Nguyen, H. N.; Tanaka, D.; Chemosphere 2022, 299, 134365. [Crossref]
Crossref...
observed a pore widening feedstock effect employing higher quantities of K2CO3 as an activating agent as a result of increasing meso- and macropore volumes. Aiming to produce adsorbent solids for postcombustion applications, we fxed a low feedstock:activating agent ratio (1:1) to prioritize the obtainment of microporous solids to capture CO2 in the low-pressure regime.

Characterizations

Characterization of the biochars via XPS (survey) revealed C 1s (285.1 eV), N 1s (400.1 eV) and O 1s (533.1 eV) signals for all the samples (Figure 2a). To investigate the nature of the nitrogen atoms incorporated into the samples, as shown in Figure 2b, high-resolution XPS spectra were obtained (Figures 2c,2d,2e,2f). Deconvolution of the signal centered at 400.1 eV (N 1s) using the software CasaXPS resulted in three signals. The signals at 398.6 and 400.1 eV were assigned to pyridinic and pyrrolic nitrogen, respectively.1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
,6363 Geng, J.-C.; Xue, D.-M.; Liu, X.-Q.; Shi, Y.-Q.; Sun, L.-B.; AIChE J. 2017, 63, 1648. [Crossref]
Crossref...
,6464 Wang, H.; Chen, C.; Chen, Y.; Wan, H.; Dong, L.; Guan, G.; J. Environ. Chem. Eng. 2021, 9, 105046. [Crossref]
Crossref...
,6565 Li, J.; Li, X.; Zhao, P.; Lei, D. Y.; Li, W.; Bai, J.; Ren, Z.; Xu, X.; Carbon 2015, 84, 460. [Crossref]
Crossref...
The signal at 402.5 eV was in the range of graphitic nitrogen (401.1-402.7 eV) and pyridinic nitrogen oxide (402.6 eV).4949 Ayiania, M.; Smith, M.; Hensley, A. J. R.; Scudiero, L.; McEwen, J. S.; Garcia-Perez, M.; Carbon 2020, 162, 528. [Crossref]
Crossref...
,6666 Kim, G.; Lee, J.; Liu, T.; Grey, C. P.; J. Phys. Chem. C 2021, 125, 10558. [Crossref]
Crossref...
Considering the anoxic environment present during the preparation of N-doped materials under a N2 atmosphere, graphitic N was more likely to form than pyridinic N-oxide was. The percentages calculated from the areas of the C 1s, N 1s and O 1s signals in the survey spectra, as well as the percentage of each type of nitrogen-containing functional group found in the high-resolution spectra, are shown in Table 1. The nitrogen percentages obtained from elemental analyses are also listed in Table 1 for comparison.

Figure 2.
(a) XPS survey spectra of Bio-500, N-Bio-600, N-Bio-700 and N-Bio-800. The Si peaks are contaminations from glass wool used in the activation process. (b) Possible N-containing functional groups in the biochars. High-resolution N 1s spectra of (c) Bio-500, (d) N-Bio-600, (e) N-Bio-700 and (f) N-Bio-800.

Table 1.
Percentages of C, O and N atoms in the biochars determined by X-ray photoelectron spectroscopy (XPS)

The percentages of nitrogen in the biochars calculated from the XPS data were similar to those obtained from the elemental analyses (Table 1). Most importantly, both nitrogen determinations (XPS and elemental analysis) showed similar tendencies involving decreases in the nitrogen percentage as the activation temperature increased.

The Raman spectra of the biochars are displayed in Figure 3a. There were two bands, one ca. 1350 cm-1 and the other ca. 1580 cm-1, related to the D-band and G-band, respectively. The G signal is due to the E2g phonon, and the D band is due to the breathing modes of sp2 rings and requires defects for activation.6767 Ferrari, A. C.; Robertson, J.; Phys. Rev. B 2000, 61, 14095. [Crossref]
Crossref...
,6868 Casiraghi, C.; Hartschuh, A.; Qian, H.; Piscanec, S.; Gerogi, C.; Fasoli, A.; Novoselov, K. S.; Basko, D. M.; Ferrari, A. C.; Nano Lett. 2009, 9, 1433. [Crossref]
Crossref...
The ID/IG ratio indicates the degree of structural defects in the biochar materials. The ID/IG values for Bio-500 and N-Bio-600 were determined to be 0.80 and 0.96, respectively, suggesting a signifcant increase in the proportion of structural defects in the latter biochar. These defects may include edges, grain boundaries, vacancies, inserted atoms, and changes in carbon hybridization (from sp2 to sp3).4848 Fattahi, M.; Taban, E.; Soltani, P.; Berardi, U.; Khavanin, A.; Zaroushani, V.; J. Building Engineering 2023, 77, 107468. [Crossref]
Crossref...
The ID/IG ratios for N-Bio-700 and N-Bio-800 were 1.00 and 1.02, respectively. By comparing the ID/IG ratios of the N-doped biochars (N-Bio-600, N-Bio-700, and N-Bio-800) with that of the precursor Bio-500, we observed a gradual increase in the number of structural defects in the N-doped samples (Table S1, Supplementary Information (SI) section). The enhanced defect formation in N-Bio-600 compared to that in Bio-500 may be attributed to the incorporation of nitrogen atoms into the carbon structure of the precursor and the activation process. However, the smoother increases in defects observed for N-Bio-600, N-Bio-700, and N-Bio-800 were most likely infuenced by the activation temperature rather than by nitrogen insertion into the carbon structures.

Figure 3.
(a) Raman spectra and (b) PXRD patterns of Bio-500, N-Bio-600, N-Bio-700 and N-Bio-800.

PDXR patterns revealed low crystallinities for all the biochars, as expected (Figure 3b). The two broad peaks at 2θ ca. 24° (002) and 43° (100) were attributed to the interlayer distances and 2D in-plane graphitic structures, respectively.6969 Lee, S. M.; Lee, S. H.; Roh, J. S.; Crystals 2021, 11, 153. [Crossref]
Crossref...
,7070 Folomeshkin, M. S.; Pisarevsky, Y. V.; Prosekov, P. A.; Volkovsky, Y. A.; Kumskov, A. S.; Grigoriev, Y. V.; Ligacheva, E. A.; Targonskii, A. V.; Blagov, A. E.; Kovalchuk, M. V.; Crystallogr. Rep. 2019, 64, 1. [Crossref]
Crossref...

N2 physisorption isotherms and cumulative pore volume curves were recorded to investigate the porosity of the biochars (Figures 4a,4b and Table 2). The specifc surface areas (SBET) obtained by the BET method were 355 m2 g-1 for Bio-500, 1095 m2 g-1 for N-Bio-600, 1521 m2 g-1 for N-Bio-700, and 1944 m2 g-1 for N-Bio-800. According to the IUPAC classifcation, all of the biochars exhibited combinations of type I and IV isotherms.7171 Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.; Rouquérol, J.; Siemieniewska, T.; Pure Appl. Chem. 1985, 57, 603. [Crossref]
Crossref...
Type I isotherms are characteristic of microporous materials, in which solidgas interactions occur predominantly at low pressures (p/p0 < 0.01). With increasing activation temperature, the saturation points of the N2 isotherm curves shifted to higher pressures, indicating enlarged pore sizes. The cumulative pore volume distributions, which were calculated using the Horvath-Kawazoe method,7272 Horvath, G.; Kawazoe, K.; J. Chem. Eng. Jpn. 1983, 16, 470. [Crossref]
Crossref...
provided insight into the volumes of pores smaller than 1 nm, between 1 and 2 nm, and larger than 2 nm (Table 2). Furthermore, we plotted the volumes of micropores smaller than 1 nm and larger than 1 nm for the biochars (Figure 4c). The results showed that the total volume of micropores smaller than 1 nm was twice the volume of pores larger than 1 nm (including micropores and mesopores). Considering the volumes for micropores smaller than 1 nm, as the activation temperature increased, the micropore volume increased linearly, but the slopes of the curves decreased as the temperature increased from N-Bio-700 to N-Bio-800 (Figure 4c). For pore volumes larger than 1 nm, which included micropores and mesopores, the opposite effect occurred, with increasing slopes observed for N-Bio-700 to N-Bio-800. Our fndings were consistent with results described in the literature.7373 Mckee, D. W.; Carbon 1982, 20, 59. [Crossref]
Crossref...
7474 Nzihou, A.; Stanmore, B.; Sharrock, P.; Energy 2013, 58, 305. [Crossref]
Crossref...
7575 Kim, M. J.; Choi, S. W.; Kim, H.; Mun, S.; Lee, K. B.; Chem. Eng. J. 2020, 397, 125404. [Crossref]
Crossref...
For example, McKee7373 Mckee, D. W.; Carbon 1982, 20, 59. [Crossref]
Crossref...
investigated the catalytic effects of various alkali metal carbonate salts on the gasifcation of carbon-based materials and observed a maximum carbon gasifcation rate when using K2CO3 at 800 °C, which increased the volume contribution from pores larger than 2 nm. Similarly, Kim et al.7575 Kim, M. J.; Choi, S. W.; Kim, H.; Mun, S.; Lee, K. B.; Chem. Eng. J. 2020, 397, 125404. [Crossref]
Crossref...
reported an increase in the mesopore volume of biochar produced from coffee grains activated with K2CO3 at temperatures above 800 °C.

Figure 4.
(a) N2 isotherms and (b) cumulative pore volume of Bio-500, N-Bio-600, N-Bio-700 and N-Bio-800. (c) Correlations between the volumes of micropores smaller than 1 nm and pores larger than 1 nm for the biochars.

Table 2.
Textural parameters, percentages of nitrogen atoms from elemental analyses, and CO2 adsorption capacities of Bio-500, N-Bio-600, N-Bio-700 and N-Bio-800 at 1 bar

STEM images of the activated N-doped biochars revealed sponge-like structures with pore sizes on the nanometer scale, whereas the precursor Bio-500 showed a less porous structure, indicating that K2CO3 effciently activated the materials (Figure 5). The images also revealed that as the activation temperature increased, the structures of the biochars became more leafed, porous, curved and wrinkled, thereby increasing the porosities of the samples, which was consistent with the BET surface areas (Table 2).

Figure 5.
STEM images of (a) and (b) Bio-500, (c) and (d) N-Bio-600, (e) and (f) N-Bio-700 and (g) and (h) N-Bio-800 obtained at different magnifications.

CO2 adsorption isotherms

The CO2 adsorption isotherms indicating the amount of CO2 adsorbed by the biochars in mmol of CO2 per gram of biochar were obtained at 25 °C and at low pressures, and the results are shown in Figure 6a and Table 2. All of the N-doped activated biochars displayed higher CO2 adsorption capacities than did the nonactivated biochar Bio-500, revealing that the addition of urea and the activation processes involving K2CO3 at different temperatures increased the CO2 adsorption capacities of the solids. This difference was related to the combination of several properties of the N-doped activated biochars, such as greater surface areas and higher micropore volumes than those of Bio-500, particularly for pores with sizes smaller than 1 nm. Additionally, the amount of nitrogen inserted in the carbon structures of the N-doped biochars may also have important effects on the CO2 adsorption capacities and selectivities (Table 3).

Figure 6.
(a) CO2 adsorption isotherms of Bio-500, N-Bio-600, N-Bio-700 and N-Bio-800 at 25 °C. (b) A plot relating the CO2 adsorption capacities at 1 bar, the narrow micropore volumes (< 1 nm) and the nitrogen contents obtained from the XPS analyses to the temperatures used for activation of the N-doped biochars.

Table 3.
Comparison between CO2 selectivities and other properties of the N-doped biochars prepared in this work and those of N-doped biochars described in the literature

N-Bio-800, with a BET surface area of 1944 m2 g-1, exhibited a lower CO2 adsorption capacity (3.70 mmol g-1) at 1 bar than did N-Bio-700 (4.23 mmol g-1), with a BET surface area of 1521 m2 g-1. It is well known from the literature that under low pressures, microporous materials are efficient gas adsorbents because the small distances between opposite walls produce effective adsorption force fields.8080 Sevilla, M.; Fuertes, A. B.; Energy Environ. Sci. 2011, 4, 1765. [Crossref]
Crossref...
We plotted a graph relating the narrow micropore volumes (< 1 nm) and nitrogen percentages to the activation temperatures of the N-doped biochars (Figure 6b). When comparing biochars Bio-500, N-Bio-600 and N-Bio-700, increases in micropore volume (< 1 nm) were followed by increased CO2 adsorption (1.66-4.73 mmol g-1). However, as previously mentioned, N-Bio-800 had a greater micropore volume and lower CO2 adsorption capacity than did N-Bio-700, indicating that other properties, such as the percentage of nitrogen, should be considered to explain this result. When the percentage of nitrogen increased to an activation temperature of 600 °C, the value decreased, e.g., the nitrogen percentage for N-Bio-800 was 0.54%, while that for N-Bio-700 was 2.60% (Table 2). The insertion of nitrogen atoms into the carbon structure can improve the CO2 adsorption capacity because nitrogen is more electronegative than carbon.1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
Nitrogen atoms produce localized charge regions, improving the interactions between CO2 and the adsorbent.1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
The results obtained in the present work indicate a synergistic effect between the micropore volume (< 1 nm) and the percentage of nitrogen inserted in the N-doped biochars, which is dependent on the activation temperature, determining the CO2 adsorption capacity. Therefore, the N-biochar activated at 700 °C had a high nitrogen percentage and high volume for micropores smaller than 1 nm, which led to the highest CO2 adsorption capacity among the synthesized materials.

Equilibrium isotherm models

Figure 7 shows the experimental CO2 adsorption isotherms measured at 25, 50 and 75 °C for Bio-500 and N-Bio-X (X = 600, 700 and 800 °C), respectively, and their respective fits to the Langmuir, Freundlich and Toth models.8181 Vargas, D. P.; Giraldo, L.; Moreno-Piraján, J. C.; Int. J. Mol. Sci. 2012, 13, 8388. [Crossref]
Crossref...
,8282 Goel, C.; Kaur, H.; Bhunia, H.; Bajpai, P. K.; J. CO2 Util. 2016, 16, 50. [Crossref]
Crossref...
The parameters obtained from each fitted equilibrium model are displayed in Table S2 (SI section).

Figure 7.
Experimental CO2 adsorption isotherms and curve fits with the Langmuir, Freundlich and Toth equations for (a) N-Bio-600, (b) N-Bio-700 and (c) N-Bio-800 measured at 25, 50 and 75 °C.

In the Langmuir model, the adsorbate remains static on the adsorbent surface, and the interactions between them present a homogeneous energy distribution.8383 Singh, V. K.; Anil Kumar, E.; Appl. Therm. Eng. 2016, 97, 77. [Crossref]
Crossref...
The Langmuir equation is written as follows:

(4) q e = q m L K L P 1 + K L P

where is the amount of CO2 adsorbed (mmol g1) at pressure P,qmL(mmol g1) is the maximum monolayer adsorption capacity and KL (bar1) is the Langmuir constant related to the free energy of adsorption. The Freundlich equation was given by He et al.:7676 He, S.; Chen, G.; Xiao, H.; Shi, G.; Ruan, C.; Ma, Y.; Dai, H.; Yuan, B.; Chen, X.; Yang, X.; J. Colloid Interface Sci. 2021, 582, 90. [Crossref]
Crossref...

(5) q e = K F P 1 n F

where KF (mmol g-1 bar-nF) and nF are parameters that refer to the adsorption capacity and adsorption intensity between the adsorbent and adsorbate, respectively.

The Toth isotherm is usually applied to describe heterogeneous adsorption systems and is expressed as follows:8484 Shafeeyan, M. S.; Daud, W. M. A. W.; Shamiri, A.; Aghamohammadi, N.; Energy Fuels 2015, 29, 6565. [Crossref]
Crossref...

(6) q e = q m T K T P ( 1 + ( K T P ) n T ) 1 n T

where P is the pressure (bar), qmT (mmol g-1) is the saturation loading, KT (bar-1) is the equilibrium constant and nT characterizes the heterogeneity of the system.

The goodness of fit between the experimental equilibrium data and the models was determined from the normalized standard deviation (∆q):

(7) Δ q = 100 × [ ( q e ( exp ) q e ( mod ) ) / q e ( exp ) ] 2 n 1

where n is the number of experimental points and the subscripts “exp” and “mod” refer to the experimental adsorption data and the adsorption data calculated by the models, respectively.8484 Shafeeyan, M. S.; Daud, W. M. A. W.; Shamiri, A.; Aghamohammadi, N.; Energy Fuels 2015, 29, 6565. [Crossref]
Crossref...

From the values of ∆q (Table S2, SI section), we conclude that the Toth isotherm model gave the best fit to the CO2 adsorption equilibrium data for the N-doped biochars. Notably, for the CO2 isotherms of N-Bio-600, N-Bio-700 and N-Bio-800, the values were less than 1 (Table S2), revealing that the biochar surfaces were energetically heterogeneous and that the CO2 was absorbed to form multilayers.8585 Chowdhury, S.; Balasubramanian, R.; Ind. Eng. Chem. Res. 2016, 55, 7906. [Crossref]
Crossref...
The equilibrium constant (KT) decreased as the adsorption temperature increased (Table S2), indicating that high adsorption temperatures disfavored solid-gas interactions. Moreover, the KT values obtained at the same adsorption temperature for N-Bio-600, N-Bio-700 and N-Bio-800 decreased as the N-doped biochar activation temperature increased, suggesting that the strength of the solid-gas interactions decreased in the order N-Bio-600 > N-Bio-700 > N-Bio-800. This difference in solid-gas interaction strength could be associated with the greater percentage of nitrogen in the N-Bio-600 sample than in the N-Bio-700 and N-Bio-800 samples. As we previously mentioned, the nitrogen atoms inserted into biochars could improve the interactions between CO2 molecules and adsorbents through the formation of localized charge regions. Most likely, based on the values (< 1) determined by the Toth model, these nitrogen atoms might be heterogeneously distributed across the surface.

The N2 isotherms measured at 25 °C for N-Bio-X (X = 600, 700 and 800 °C) were also fitted with the Langmuir, Freundlich and Toth models (Figure S1 and Table S3, SI section). The Toth model again provided the best fit. In contrast to CO2 adsorption, the N2 adsorption capacity was independent of the nitrogen percentage present in the N-doped biochars. Moreover, the N2 adsorption capacity increased as the specific surface area increased.

Isosteric heats of CO2 adsorption

The Clausius-Clapeyron equation (equation 8) was used to calculate the isosteric heats of CO2 adsorption on the N-doped biochars, with the adsorption isotherms measured at 25, 50 and 75 °C (Figure S2). Due to the goodness of fit between the Toth model and the experimental equilibrium data, we used such curves to apply the Clausius-Clapeyron equation.

(8) Q st = R [ ln ( P ) ( 1 / T ) ] n

where Qst (kJ mol-1) is the isosteric heat for the amount adsorbed n (mmol g-1), P (bar) is the pressure, T (K) is the adsorption temperature and R (8.314 10-3 kJ mol K-1) is the universal gas constant. The slope of the ln (P) versus 1/T plot was ln(P)(1/T) (Figure S3). The negative values of the isosteric heats (Qst) observed in Figure 8 confirmed the exothermic nature of the process and explained the decreased CO2 adsorption capacities of the biochars with increases in the temperatures of the isotherms, as observed in Figure 7. The isosteric heat decreased for all the N-doped biochars as the CO2 coverage increased from 0.01 to 2.00 mmol g1 (Figure 8). Most likely, the CO2 molecules preferentially adsorbed on the most energetic sites on the surfaces of the biochars, i.e., the nitrogen sites, and then, as these sites were occupied, adsorption occurred on the less energetic sites.8686 Musa, S. G.; Aljunid Merican, Z. M.; Haruna, A.; J. Solid State Chem. 2022, 314, 123363. [Crossref]
Crossref...
Moreover, the greater surface heterogeneity of N-Bio-600 than of the other samples was indicated by the most prominent variation in its isosteric heat as the CO2 coverage increased. Considering the uptake of 0.01 mmol g-1 CO2 (Figure 8), the calculated isosteric heat became less negative as the activation temperature of the N-doped biochar increased, e.g., -34.15 kJ mol-1 for N-Bio-600, -30.50 kJ mol-1 for N-Bio-700 and -27.43 kJ mol-1 for N-Bio-800. These results were consistent with those obtained from the Toth model, which confrmed that the N-Bio-600 sample, which had the highest nitrogen percentage and was the most energetic and heterogeneous surface, formed the strongest interactions with CO2. In addition, the type of nitrogen incorporated in biochar, e.g., pyridinic, graphitic or pyrrolic, is important for solid-gas interactions.1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
All nitrogen species may interact with CO2; however, the interaction with pyridinic nitrogen might be stronger than that with pyrrolic nitrogen because the lone pair on the nitrogen atom of the pyridinic ring is available to interact with the acidic CO2 molecule. The graphitic nitrogen types, particularly those containing nitrogen in the meta position of the ring, can strongly interact with CO2 due to the high polarization caused by nitrogen substitution. Indeed, N-Bio-600, which presented the most negative isosteric heat (Qst), was the solid containing the highest percentage of pyridinic and graphitic nitrogen atoms (0.99%) in comparison with N-Bio-700 (0.72%) and N-Bio-800 (0.13%) (see Table 1).

Figure 8.
Isosteric heats for CO2 adsorption onto N-Bio-600, N-Bio-700 and N-Bio-800.

The relatively low values are commonly associated with the physisorption process. According to the literature, the optimal range is typically -30 to -50 kJ mol-1. Materials with higher isosteric heats of adsorption require more energy to undo the formed interactions and promote desorption, which can have negative economic implications.8787 Patel, H. A.; Byun, J.; Yavuz, C. T.; ChemSusChem 2017, 10, 1303. [Crossref]
Crossref...

Selectivity studies (CO2/N2)

Postcombustion is used in thermoelectric power plants and in the fertilizer industry. The combustion of coal or natural gas in the presence of air releases a gas fow at a pressure of 1 bar, which contains a low concentration of CO2 (approximately 15%), a high content of N2 (approximately 78%) and low percentages of other gases such as H2O, O2, CO, NOx and SOx.1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
Therefore, the separation between CO2 and N2 must occur via adsorbents with high CO2 selectivities at low pressures. The recovery of adsorbents containing CO2 is conducted by heating and/or under vacuum.1313 dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
Crossref...
Therefore, it is necessary to investigate the selectivities of new adsorbent materials for CO2 and N2.

Figure 9a shows the CO2 and N2 adsorption branch isotherms measured at 25 °C. All of the N-doped biochars adsorbed much greater amounts of CO2 than N2. Because CO2 has a smaller kinetic diameter, a higher quadrupole moment (CO2, 13.4 × 10-40 C m2; N2, 4.7 × 10-40 C m2) and a higher polarizability (CO2, 29.0 × 10-25 cm-3; N2, 17.4 × 10-25 cm-3) than N2, more CO2 is expected to be adsorbed on N-doped biochars.7171 Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.; Rouquérol, J.; Siemieniewska, T.; Pure Appl. Chem. 1985, 57, 603. [Crossref]
Crossref...

Figure 9.
(a) CO2 and N2 adsorption isotherms measured at 25 °C (the open symbols refer to N2, and the closed symbols refer to CO2) and (b) the IAST CO2 selectivities of N-Bio-600, N-Bio-700 and N-Bio-800.

A comparison of the N2 adsorption capacities of the N-Bio-X materials (X = 600, 700 and 800 °C) revealed that the increase in adsorption strength was proportional to the increase in activation temperature. This result is ultimately related to the increased specific surface areas and pore volumes of N-Bio-X (Table 2). In contrast, the CO2 adsorption behavior did not depend only on the increased surface areas of the biochars, as previously discussed.

The selectivity for a binary mixture of CO2 and N2 at 25 °C can be calculated with the ideal adsorbed solution theory (IAST) from the single-component adsorption isotherms for CO2 and N2.7272 Horvath, G.; Kawazoe, K.; J. Chem. Eng. Jpn. 1983, 16, 470. [Crossref]
Crossref...
For the selectivity calculation, the ratio in the nonadsorbed phase was 85/15, which is the typical composition of the gases exhausted in the postcombustion process for a binary composition.7373 Mckee, D. W.; Carbon 1982, 20, 59. [Crossref]
Crossref...
,7474 Nzihou, A.; Stanmore, B.; Sharrock, P.; Energy 2013, 58, 305. [Crossref]
Crossref...
The selectivity can be calculated by solving equations 9 and 10 for the variable .8282 Goel, C.; Kaur, H.; Bhunia, H.; Bajpai, P. K.; J. CO2 Util. 2016, 16, 50. [Crossref]
Crossref...
The calculation was conducted using the functions integrate and solved in the free software Scilab.5050 Scilab, 6.10; French Institute for Research in Computer Science and Control, Vélizy-Villacoublay, France, 1984. The results are usually represented in the literature with equation 11.

(9) 0 Py CO 2 x CO 2 n CO 2 ( P CO 2 ) P CO 2 dP CO 2 = 0 Py N 2 x N 2 n N 2 ( P N 2 ) P N 2 dP N 2

(10) x N 2 + x CO 2 = 1

(11) IAST CO 2 / N 2 = x CO 2 x N 2 y N 2 y CO 2

where P is the pressure, PCO2 and PN2 are the partial pressures of CO2 and N2, xCO2 and xN2 are the molar fractions of CO2 and N2 in the adsorbed phase, yCO2 and yN2 are the molar fractions of CO2 and N2 in the nonadsorbed phase, and nCO2 and nN2 can be represented by adsorption equilibrium models; for the reasons already presented, we decided to use the Toth model. Figure 9b shows the IAST selectivities.

The N-doped biochars studied in this work exhibited xCO2/xN2>1, considering yN2/yCO2=85/15. Therefore, all the N-doped biochars were more selective for CO2 than for N2, and N-Bio-600 was the most selective. Although N-Bio-700 adsorbs more CO2 than does N-Bio-600, the selectivity of the former was 2.8 times lower than that of the latter.

Table 3 shows a comparison of the CO2 selectivities, nitrogen percentages, BET surface areas and CO2 adsorption capacities at 1 bar for the N-doped biochars prepared in this work with those of N-doped biochars obtained from different biomasses and described in the literature.

Table 3 shows that N-Bio-600 was the most selective toward CO2, but N-Bio-700 adsorbed more CO2. As previously discussed, increasing the activation temperature of the biochar resulted in increases in the specific surface area and total pore volume, which were beneficial for N2 adsorption. However, a high activation temperature enlarged the pore sizes (Table 3) and reduced the nitrogen percentage on the surfaces of the biochars. Consequently, the selectivity toward CO2 decreased because the larger pores also adsorbed N2, which presented a larger kinetic diameter. Additionally, the smaller the nitrogen percentage in the sample was, the weaker the interaction with CO2 was, which decreased the selectivity.

Kinetic studies

Figure 10a shows the results of the CO2 adsorption kinetic studies. The kinetic behaviors of the N-Bio-X (X = 600, 700 and 800 °C) samples were similar to those of Bio-500. High adsorption rates were observed in the first 45 s. After one minute of the experiment, the adsorption rate decreased until kinetic equilibrium was reached. After three minutes, no significant changes in CO2 adsorption were observed. There are reports of much longer times (ca. 60 min) required to reach CO2 adsorption kinetic equilibrium for N-carbon-based materials.7575 Kim, M. J.; Choi, S. W.; Kim, H.; Mun, S.; Lee, K. B.; Chem. Eng. J. 2020, 397, 125404. [Crossref]
Crossref...
,8888 Yuan, X.; Li, S.; Jeon, S.; Deng, S.; Zhao, L.; Lee, K. B.; J. Hazard Mater. 2020, 399, 123010. [Crossref]
Crossref...
All the materials prepared in this work exhibited similar adsorption rates, i.e., 0.45, 0.38 and 0.46 mmol g-1 min-1 for N-Bio-600, N-Bio-700 and N-Bio-800, respectively, demonstrating fast adsorption and diffusion of CO2 in the pores. This result agrees with the low isosteric heat values (Figure 8) typical of physical adsorption driven by van der Waals forces, which does not involve activation energy, resulting in a fast adsorption process.8989 Chang, Q.; Colloid and Interface Chemistry for Water Quality Control; Elsevier, 2016, ch. 10. [Crossref]
Crossref...
Adsorbents with a combination of micro- and mesopores show good mass transfer.9090 Mukhtar, A.; Saqib, S.; Mellon, N. B.; Babar, M.; Rafiq, S.; Ullah, S.; Bustam, M. A.; Al-Sehemi, A. G.; Muhammad, N.; Chawla, M.; J. Nat. Gas Sci. Eng. 2020, 77, 103203. [Crossref]
Crossref...
The N-doped biochars prepared in this work presented comparable mesopore volume percentages: 9.42% (N-Bio-600), 8.42% (N-Bio-700) and 10.0% (N-Bio-800). Because the interactions between CO2 and mesopores are energetically less favorable than those between CO2 and micropores, the presence of mesopores will facilitate gas access to micropores. Therefore, a higher mesopore volume is expected to result in faster adsorption. As the N-doped biochars prepared in this work exhibited similar mesopore volume percentages, the adsorption rates were also similar. Adsorbents with fast adsorption rates are interesting for practical carbon capture applications due to the high exhaust gas flow rates observed in industry.

Figure 10.
(a) CO2 adsorption kinetic curves for Bio-500, N-Bio-600, N-Bio-700 and N-Bio-800 and (b) CO2 adsorption and desorption cycles for N-Bio-700 measured at 30 °C (the desorption time was the same for each cycle).

Considering that more CO2 was adsorbed by N-Bio-700 than by the other synthesized biochars, we investigated cyclic CO2 adsorption to evaluate its reusability (Figure 10b). The results revealed that N-Bio-700 remained stable for up to seven CO2 adsorption/desorption cycles. Desorption was induced by flowing nitrogen at room temperature, possibly due to the low isosteric heat of N-Bio-700. This characteristic is important since, in chemical absorption technologies, the heat required for the regeneration of amine aqueous solutions makes the process very costly.9191 Aghel, B.; Janati, S.; Wongwises, S.; Shadloo, M. S.; Int. J. Greenhouse Gas Control 2022, 119, 103715. [Crossref]
Crossref...

Conclusions

In conclusion, microporous N-doped activated biochars derived from corn husks, K2CO3 as the activating agent and urea as the doping compound showed promise for CO2 capture in postcombustion processes. At 25 °C and 1 bar, CO2 adsorption studies showed that the adsorption capacities increased in the order Bio-500 (1.66 mmol g-1) < N-Bio-600 (3.38 mmol g-1) < N-Bio-800 (3.70 mmol g-1) < N-Bio-700 (4.23 mmol g-1). N-Bio-700 exhibited the highest performance due to the synergistic combination of the volume derived from micropores smaller than 1 nm (0.51 cm3 g-1) and the percentage of nitrogen atoms (2.60%) on the surface. The Toth model describes the CO2 and N2 isotherms for all the N-doped biochars, and from the fitting parameters, the isosteric adsorption heat and the selectivity were calculated for a binary mixture of N2/CO2 (85/15). The isosteric heats ranged from -25 to -34 kJ mol-1, confirming that physisorption was the predominant adsorption mode for CO2 capture. Moreover, N-Bio-600 was the most energetic and heterogeneous surface because it had the highest percentage of nitrogen (5.02%). This material also demonstrated the highest selectivity toward CO2 under the studied conditions. Kinetic studies revealed that adsorption equilibrium was reached in less than 3 min with the biochars. Regeneration studies demonstrated that N-Bio-700 maintained stability over multiple adsorption and desorption cycles. Therefore, both N-Bio-600 and N-Bio-700 are very promising biochars for CO2 capture at low pressures.

Acknowledgments

This work was supported by the Brazilian agencies Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) and Fundação Carlos Chagas Filho de Amparo à Pesquisa do Estado do Rio de Janeiro (FAPERJ). We are grateful to Molecular Spectroscopy and Material Characterization Multiuser Laboratories from Fluminense Federal University.

Supplementary Information

Supplementary information (Raman wavenumber and ID/IG ratio for Bio-500 and N-Bio-X; fitted parameters of the experimental CO2 isotherms of N-Bio-X; experimental and curve fittings of N2 isotherms; fitted parameters of the experimental N2 isotherms of N-Bio-X; CO2 adsorption isotherms of N-Bio-X; and ln (P) versus 1/T (K-1) plots used to calculate the isosteric heat of adsorption (Qst) for N-Bio-X) is available free of charge at http://jbcs.sbq.org.br as PDF file.

  • Dedicated to the memory of Prof Oswaldo Luiz Alves for his contribution to the feld of solid-state chemistry and for his passion for and dedication to science and teaching.

References

  • 1
    Pachauri, R. K.; Allen, M. R.; Barros, V. R.; Broome, J.; Cramer, W.; Christ, R.; Church, J. A.; Clarke, L.; Dahe, Q.; Dasgupta, P.; Dubash, N. K.; Edenhofer, O.; Elgizouli, I.; Field, C. B.; Forster, P.; Friedlingstein, P.; Fuglestvedt, J.; Gomez-Echeverri, L.; Hallegatte, S.; Hegerl, G.; Howden, M.; Jiang, K.; Cisneros, B. J.; Kattsov, V.; Lee, H.; Mach, K. J.; Marotzke, J.; Mastrandrea, M. D.; Meyer, L.; Minx, J.; Mulugetta, Y.; O’Brien, K.; Oppenheimer, M.; Pereira, J. J.; Pichs-Madruga, R.; Plattner, G.-K.; Pörtner, H.-O.; Power, S. B.; Preston, B.; Ravindranath, N. H.; Reisinger, A.; Riahi, K.; Rusticucci, M.; Scholes, R.; Seyboth, K.; Sokona, Y.; Stavins, R.; Stocker, T. F.; Tschakert, P.; van Vuuren, D.; van Ypersele, J.-P.; Climate Change 2014: Mitigation of Climate Change, IPCC Fifth Assessment Report, Cambridge University Press, 2015. [Crossref]
    » Crossref
  • 2
    Global CCS Institute, Global Status of CCS Report: 2020, https://www.globalccsinstitute.com/resources/publications-reports-research/global-status-of-ccs-report-2020/, accessed in February 2024.
    » https://www.globalccsinstitute.com/resources/publications-reports-research/global-status-of-ccs-report-2020/
  • 3
    Chagas, J. A. O.; Marciniak, A. A.; Mota, C. J. A.; J. Braz. Chem. Soc. 2022, 33, 801. [Crossref]
    » Crossref
  • 4
    dos Santos, T. C.; Lage, M. R.; da Silva, A. F. M.; Fernandes, T. S.; Carneiro, J. W. M.; Ronconi, C. M.; J. CO2 Util. 2022, 61, 102054. [Crossref]
    » Crossref
  • 5
    Furtado, I. O.; dos Santos, T. C.; Vasconcelos, L. F.; Costa, L. T.; Fiorot, R. G.; Ronconi, C. M.; Carneiro, J. W. M.; Chem. Eng. J. 2021, 408, 128002. [Crossref]
    » Crossref
  • 6
    de Vasconcelos, S. F.; Carneiro, L. O.; Brito, R. P.; Brito, K. D.; J. Braz. Chem. Soc. 2023, 34, 441. [Crossref]
    » Crossref
  • 7
    Aquino, A. S.; Bernard, F. L.; Vieira, M. O.; Borges, J. V.; Rojas, M. F.; Vecchia, F. D.; Ligabue, R. A.; Seferin, M.; Menezes, S.; Einloft, S.; J. Braz. Chem. Soc. 2014, 25, 2251. [Crossref]
    » Crossref
  • 8
    Freitas, S. K. S.; Oliveira, F. L.; dos Santos, T. C.; Hisse, D.; Merlini, C.; Ronconi, C. M.; Esteves, P. M.; Chem. Eur. J. 2021, 27, 2342. [Crossref]
    » Crossref
  • 9
    dos Santos, T. C.; Ronconi, C. M.; J. CO2 Util. 2017, 20, 292. [Crossref]
    » Crossref
  • 10
    Wang, X.; Song, C.; Front. Energy Res. 2020, 8, 265. [Crossref]
    » Crossref
  • 11
    Quang, D. V.; Milani, D.; Abu Zahra, M.; Int. J. Greenhouse Gas Control 2023, 124, 103862. [Crossref]
    » Crossref
  • 12
    Ribeiro, M. G.; Hisse, D.; Prado, M. L.; dos Santos, T. C.; Ronconi, C. M.; Rev. Virtual Quim. 2022, 14, 517. [Crossref]
    » Crossref
  • 13
    dos Santos, T. C.; Mancera, R. C.; Rocha, M. V. J.; da Silva, A. F. M.; Furtado, I. O.; Barreto, J.; Stavale, F.; Archanjo, B. S.; De Carneiro, J. W. M.; Costa, L. T.; Ronconi, C. M.; J. CO2 Util. 2021, 48, 101517. [Crossref]
    » Crossref
  • 14
    dos Santos, T. C.; Bourrelly, S.; Llewellyn, P. L.; Carneiro, J. W. D. M.; Ronconi, C. M.; Phys. Chem. Chem. Phys. 2015, 17, 11095. [Crossref]
    » Crossref
  • 15
    Mello, M. R.; Phanon, D.; Silveira, G. Q.; Llewellyn, P. L.; Ronconi, C. M.; Microporous Mesoporous Mater. 2011, 143, 174. [Crossref]
    » Crossref
  • 16
    dos Santos, T. C.; Ronconi, C. M.; Rev. Virtual Quim. 2014, 6, 112. [Crossref]
    » Crossref
  • 17
    Babu, P. E.; Tejes, P. K. S.; Naik, B. K.; Carbon Capture Sci. Technol. 2023, 7, 100103. [Crossref]
    » Crossref
  • 18
    Najafi, A. M.; Soltanali, S.; Khorashe, F.; Ghassabzadeh, H.; Chemosphere 2023, 324, 138275. [Crossref]
    » Crossref
  • 19
    Dinda, S.; Mater. Today Commun. 2023, 35, 105927. [Crossref]
    » Crossref
  • 20
    Salih, H. A.; Alkhatib, I. I. I.; Zahra, M. A.; Vega, L. F.; J. CO2 Util. 2023, 68, 102383. [Crossref]
    » Crossref
  • 21
    Short, G. N.; Burentugs, E.; Proaño, L.; Moon, H. J.; Rim, G.; Nezam, I.; Korde, A.; Nair, S.; Jones, C. W.; JACS Au 2023, 3, 62. [Crossref]
    » Crossref
  • 22
    Huang, Z.; Ying, L.; Gong, F.; Lu, J.; Wang, W.; Ding, J.; Yan, J.; J. Environ. Chem. Eng. 2023, 11, 109739. [Crossref]
    » Crossref
  • 23
    Ahn, H.; Xu, R.; Kim, S.; Lee, J. W.; Kang, Y. T.; J. Environ. Chem. Eng. 2023, 11, 109723. [Crossref]
    » Crossref
  • 24
    Zhang, X.; Cao, L.; Xiang, W.; Xu, Y.; Gao, B.; Sep. Purif. Technol. 2022, 295, 121295. [Crossref]
    » Crossref
  • 25
    Bamdad, H.; Hawboldt, K.; Macquarrie, S.; Energy Fuels 2018, 32, 11742. [Crossref]
    » Crossref
  • 26
    Chatterjee, R.; Sajjadi, B.; Chen, W. Y.; Mattern, D. L.; Egiebor, N. O.; Hammer, N.; Raman, V.; Energy Fuels 2019, 33, 2366. [Crossref]
    » Crossref
  • 27
    Wu, B.; Song, X.; Zheng, D.; Tan, Q.; Yao, Y.; Liu, F. Q.; ACS Appl. Mater. Interfaces 2023, 15, 20325. [Crossref]
    » Crossref
  • 28
    Li, D.; Chen, W.; Wu, J.; Jia, C. Q.; Jiang, X.; J. Mater. Chem. A 2020, 8, 24977. [Crossref]
    » Crossref
  • 29
    Kortunov, P. V.; Siskin, M.; Paccagnini, M.; Thomann, H.; Energy Fuels 2016, 30, 1223. [Crossref]
    » Crossref
  • 30
    Li, W.; Tu, W.; Cheng, J.; Yang, F.; Wang, X.; Li, L.; Shang, D.; Zhou, X.; Yu, C.; Yuan, A.; Pan, J.; Sep. Purif. Technol. 2022, 282, 120001. [Crossref]
    » Crossref
  • 31
    Li, D.; Ma, T.; Zhang, R.; Tian, Y.; Qiao, Y.; Fuel 2015, 139, 68. [Crossref]
    » Crossref
  • 32
    Yue, L.; Xia, Q.; Wang, L.; Wang, L.; DaCosta, H.; Yang, J.; Hu, X.; J. Colloid Interface Sci. 2018, 511, 259. [Crossref]
    » Crossref
  • 33
    Liang, W.; Luo, Z.; Zhou, D.; Liu, Z.; Wei, X.; Cai, W.; Chem. Eng. J. 2023, 461, 141980. [Crossref]
    » Crossref
  • 34
    Luo, Z.; Peng, X.; Liang, W.; Zhou, D.; Dang, C.; Cai, W.; Bioresour. Technol. 2023, 388, 129762. [Crossref]
    » Crossref
  • 35
    Peng, X.; Luo, Z.; Xie, H.; Liang, W.; Luo, J.; Dang, C.; Wang, A.; Hu, L.; Yu, X.; Cai, W.; Appl. Surf. Sci. 2022, 606, 154859. [Crossref]
    » Crossref
  • 36
    Bai, J.; Huang, J.; Yu, Q.; Demir, M.; Kilic, M.; Altay, B. N.; Hu, X.; Wang, L.; Fuel Process Technol. 2023, 249, 107854. [Crossref]
    » Crossref
  • 37
    Khan, M. H.; Akash, N. M.; Akter, S.; Rukh, M.; Nzediegwu, C.; Islam, M. S.; J. Environ. Manage. 2023, 338, 117825. [Crossref]
    » Crossref
  • 38
    Hanif, A.; Aziz, M. A.; Helal, A.; Abdelnaby, M. M.; Khan, A.; Theravalappil, R.; Khan, M. Y.; ACS Omega 2023, 8, 36228. [Crossref]
    » Crossref
  • 39
    Ghanbarpour Mamaghani, Z.; Hawboldt, K. A.; MacQuarrie, S.; J. Environ. Chem. Eng. 2023, 11, 109643. [Crossref]
    » Crossref
  • 40
    Zaker, A.; ben Hammouda, S.; Sun, J.; Wang, X.; Li, X.; Chen, Z.; J. Environ. Chem. Eng. 2023, 11, 109741. [Crossref]
    » Crossref
  • 41
    Adio, S. O.; Ganiyu, S. A.; Usman, M.; Abdulazeez, I.; Alhooshani, K.; Chem. Eng. J. 2020, 382, 122964. [Crossref]
    » Crossref
  • 42
    Ma, X.; Su, C.; Liu, B.; Wu, Q.; Zhou, K.; Zeng, Z.; Li, L.; Sep. Purif. Technol. 2021, 259, 118065. [Crossref]
    » Crossref
  • 43
    Rao, L.; Liu, S.; Wang, L.; Ma, C.; Wu, J.; An, L.; Hu, X.; Chem. Eng. J 2019, 359, 428. [Crossref]
    » Crossref
  • 44
    Luo, J.; Liu, B.; Shi, R.; Guo, Y.; Feng, Q.; Liu, Z.; Li, L.; Norinaga, K.; Microporous Mesoporous Mater. 2021, 327, 111404. [Crossref]
    » Crossref
  • 45
    Chun, S.-E.; Whitacre, J. F.; Microporous Mesoporous Mater. 2017, 251, 34. [Crossref]
    » Crossref
  • 46
    Hayashi, J.; Yamamoto, N.; Horikawa, T.; Muroyama, K.; Gomes, V. G.; J. Colloid Interface Sci. 2005, 281, 437. [Crossref]
    » Crossref
  • 47
    Ratna, A. S.; Ghosh, A.; Mukhopadhyay, S.; J. Cleaner Prod. 2022, 371, 133563. [Crossref]
    » Crossref
  • 48
    Fattahi, M.; Taban, E.; Soltani, P.; Berardi, U.; Khavanin, A.; Zaroushani, V.; J. Building Engineering 2023, 77, 107468. [Crossref]
    » Crossref
  • 49
    Ayiania, M.; Smith, M.; Hensley, A. J. R.; Scudiero, L.; McEwen, J. S.; Garcia-Perez, M.; Carbon 2020, 162, 528. [Crossref]
    » Crossref
  • 50
    Scilab, 6.10; French Institute for Research in Computer Science and Control, Vélizy-Villacoublay, France, 1984.
  • 51
    Cessford, N. F.; Seaton, N. A.; Düren, T.; Ind. Eng. Chem. Res. 2012, 51, 4911. [Crossref]
    » Crossref
  • 52
    Yang, J.; Yue, L.; Hu, X.; Wang, L.; Zhao, Y.; Lin, Y.; Sun, Y.; DaCosta, H.; Guo, L.; Energy Fuels 2017, 31, 4287. [Crossref]
    » Crossref
  • 53
    Yang, X.; Han, F.; Xu, C.; Jiang, S.; Huang, L.; Liu, L.; Xia, Z.; Ind. Crops Prod. 2017, 109, 241. [Crossref]
    » Crossref
  • 54
    Del Río, J. C.; Rencoret, J.; Gutiérrez, A.; Kim, H.; Ralph, J.; J. Agric. Food Chem. 2018, 66, 4402. [Crossref]
    » Crossref
  • 55
    Yeasmin, M. S.; Mondal, M. I. H.; Int. J. Biol. Macromol. 2015, 80, 725. [Crossref]
    » Crossref
  • 56
    Singh, G.; Kim, I. Y.; Lakhi, K. S.; Srivastava, P.; Naidu, R.; Vinu, A.; Carbon 2017, 116, 448. [Crossref]
    » Crossref
  • 57
    Zhu, N.; Qian, F.; Xu, X.; Wang, M.; Teng, Q.; Materials 2021, 14, 6190. [Crossref]
    » Crossref
  • 58
    Tischer, S.; Börnhorst, M.; Amsler, J.; Schoch, G.; Deutschmann, O.; Phys. Chem. Chem. Phys. 2019, 21, 16785. [Crossref]
    » Crossref
  • 59
    Bai, J.; Huang, J.; Jiang, Q.; Jiang, W.; Demir, M.; Kilic, M.; Altay, B. N.; Wang, L.; Hu, X.; Colloids Surf., A 2023, 674, 131916. [Crossref]
    » Crossref
  • 60
    Bai, J.; Huang, J.; Yu, Q.; Demi, M.; Akgul, E.; Altay, B. N.; Hu, X.; Wang, L.; Front Chem. Sci. Eng. 2023, 17, 1122. [Crossref]
    » Crossref
  • 61
    Bai, J.; Shao, J.; Yu, Q.; Demir, M.; Altay, B. N.; Ali, T. M.; Jiang, Y.; F, L. W.; Hu, X.; Chem. Eng. J. 2024, 479, 147667. [Crossref]
    » Crossref
  • 62
    Khuong, D. A.; Trinh, K. T.; Nakaoka, Y.; Tsubota, T.; Tashima, D.; Nguyen, H. N.; Tanaka, D.; Chemosphere 2022, 299, 134365. [Crossref]
    » Crossref
  • 63
    Geng, J.-C.; Xue, D.-M.; Liu, X.-Q.; Shi, Y.-Q.; Sun, L.-B.; AIChE J. 2017, 63, 1648. [Crossref]
    » Crossref
  • 64
    Wang, H.; Chen, C.; Chen, Y.; Wan, H.; Dong, L.; Guan, G.; J. Environ. Chem. Eng. 2021, 9, 105046. [Crossref]
    » Crossref
  • 65
    Li, J.; Li, X.; Zhao, P.; Lei, D. Y.; Li, W.; Bai, J.; Ren, Z.; Xu, X.; Carbon 2015, 84, 460. [Crossref]
    » Crossref
  • 66
    Kim, G.; Lee, J.; Liu, T.; Grey, C. P.; J. Phys. Chem. C 2021, 125, 10558. [Crossref]
    » Crossref
  • 67
    Ferrari, A. C.; Robertson, J.; Phys. Rev. B 2000, 61, 14095. [Crossref]
    » Crossref
  • 68
    Casiraghi, C.; Hartschuh, A.; Qian, H.; Piscanec, S.; Gerogi, C.; Fasoli, A.; Novoselov, K. S.; Basko, D. M.; Ferrari, A. C.; Nano Lett. 2009, 9, 1433. [Crossref]
    » Crossref
  • 69
    Lee, S. M.; Lee, S. H.; Roh, J. S.; Crystals 2021, 11, 153. [Crossref]
    » Crossref
  • 70
    Folomeshkin, M. S.; Pisarevsky, Y. V.; Prosekov, P. A.; Volkovsky, Y. A.; Kumskov, A. S.; Grigoriev, Y. V.; Ligacheva, E. A.; Targonskii, A. V.; Blagov, A. E.; Kovalchuk, M. V.; Crystallogr. Rep. 2019, 64, 1. [Crossref]
    » Crossref
  • 71
    Sing, K. S. W.; Everett, D. H.; Haul, R. A. W.; Moscou, L.; Pierotti, R. A.; Rouquérol, J.; Siemieniewska, T.; Pure Appl. Chem. 1985, 57, 603. [Crossref]
    » Crossref
  • 72
    Horvath, G.; Kawazoe, K.; J. Chem. Eng. Jpn. 1983, 16, 470. [Crossref]
    » Crossref
  • 73
    Mckee, D. W.; Carbon 1982, 20, 59. [Crossref]
    » Crossref
  • 74
    Nzihou, A.; Stanmore, B.; Sharrock, P.; Energy 2013, 58, 305. [Crossref]
    » Crossref
  • 75
    Kim, M. J.; Choi, S. W.; Kim, H.; Mun, S.; Lee, K. B.; Chem. Eng. J. 2020, 397, 125404. [Crossref]
    » Crossref
  • 76
    He, S.; Chen, G.; Xiao, H.; Shi, G.; Ruan, C.; Ma, Y.; Dai, H.; Yuan, B.; Chen, X.; Yang, X.; J. Colloid Interface Sci. 2021, 582, 90. [Crossref]
    » Crossref
  • 77
    Li, H.; Tang, M.; Huang, X.; Wang, L.; Liu, Q.; Lu, S.; Chem. Eng. J. 2023, 466, 143095. [Crossref]
    » Crossref
  • 78
    Wu, R.; Ye, Q.; Wu, K.; Wang, L.; Dai, H.; J. CO2 Util. 2021, 51, 101620. [Crossref]
    » Crossref
  • 79
    Gan, F.; Wang, B.; Guo, J.; He, J.; Ma, S.; Jiang, X.; Jin, Z.; Sep. Purif. Technol. 2022, 302, 122089. [Crossref]
    » Crossref
  • 80
    Sevilla, M.; Fuertes, A. B.; Energy Environ. Sci. 2011, 4, 1765. [Crossref]
    » Crossref
  • 81
    Vargas, D. P.; Giraldo, L.; Moreno-Piraján, J. C.; Int. J. Mol. Sci. 2012, 13, 8388. [Crossref]
    » Crossref
  • 82
    Goel, C.; Kaur, H.; Bhunia, H.; Bajpai, P. K.; J. CO2 Util. 2016, 16, 50. [Crossref]
    » Crossref
  • 83
    Singh, V. K.; Anil Kumar, E.; Appl. Therm. Eng. 2016, 97, 77. [Crossref]
    » Crossref
  • 84
    Shafeeyan, M. S.; Daud, W. M. A. W.; Shamiri, A.; Aghamohammadi, N.; Energy Fuels 2015, 29, 6565. [Crossref]
    » Crossref
  • 85
    Chowdhury, S.; Balasubramanian, R.; Ind. Eng. Chem. Res. 2016, 55, 7906. [Crossref]
    » Crossref
  • 86
    Musa, S. G.; Aljunid Merican, Z. M.; Haruna, A.; J. Solid State Chem. 2022, 314, 123363. [Crossref]
    » Crossref
  • 87
    Patel, H. A.; Byun, J.; Yavuz, C. T.; ChemSusChem 2017, 10, 1303. [Crossref]
    » Crossref
  • 88
    Yuan, X.; Li, S.; Jeon, S.; Deng, S.; Zhao, L.; Lee, K. B.; J. Hazard Mater. 2020, 399, 123010. [Crossref]
    » Crossref
  • 89
    Chang, Q.; Colloid and Interface Chemistry for Water Quality Control; Elsevier, 2016, ch. 10. [Crossref]
    » Crossref
  • 90
    Mukhtar, A.; Saqib, S.; Mellon, N. B.; Babar, M.; Rafiq, S.; Ullah, S.; Bustam, M. A.; Al-Sehemi, A. G.; Muhammad, N.; Chawla, M.; J. Nat. Gas Sci. Eng. 2020, 77, 103203. [Crossref]
    » Crossref
  • 91
    Aghel, B.; Janati, S.; Wongwises, S.; Shadloo, M. S.; Int. J. Greenhouse Gas Control 2022, 119, 103715. [Crossref]
    » Crossref

Edited by

Editor handled this article: Aldo José Gorgatti Zarbin (Guest)

Publication Dates

  • Publication in this collection
    19 Apr 2024
  • Date of issue
    2024

History

  • Received
    14 Dec 2023
  • Accepted
    18 Mar 2024
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