Next Article in Journal
Interface-Sensitive Charge Storage and Activation Behavior of Mn(1,3,5-Benzenetricarboxylic Acid (BTC))-Derived Mn3O4/Carbon Cathodes for Aqueous Zinc-Ion Batteries
Previous Article in Journal
Natural Bioactive Compound-Integrated Nanomaterials for Diabetic Wound Healing: Synergistic Effects, Multifunctional Designs, and Challenges
Previous Article in Special Issue
Progressive Conversion Model Applied to the Physical Activation of Activated Carbon from Palm Kernel Shells at the Pilot Scale in a Nichols Furnace and at the Industrial Scale in a Rotary Kiln
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Synthesis of B-Doped Porous Carbons via a Sodium Metaborate Tetrahydrate Activating Agent: A Novel Approach for CO2 Adsorption

1
Key Laboratory of the Ministry of Education for Advanced Catalysis Materials, Zhejiang Normal University, Jinhua 321004, China
2
Institute of Plant Nutrition, Resources and Environment, Beijing Academy of Agriculture and Forestry Science, Beijing 100097, China
3
Department of Chemical Engineering, Bogazici University, Istanbul 34342, Türkiye
4
TUBITAK Marmara Research Center, Material Institute, Gebze 41470, Türkiye
5
Chemical Engineering Department, College of Engineering, K. S. A & Central Labs, King Khalid University, Abha 61413, Saudi Arabia
6
Department of Biomedical Engineering, Faculty of Engineering and Natural Sciences, Iskenderun Technical University, Hatay 31200, Türkiye
7
Key Laboratory of Urban Rail Transit Intelligent Operation and Maintenance Technology and Equipment of Zhejiang Province, College of Engineering, Zhejiang Normal University, Jinhua 321004, China
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(12), 2564; https://doi.org/10.3390/molecules30122564
Submission received: 14 May 2025 / Revised: 9 June 2025 / Accepted: 10 June 2025 / Published: 12 June 2025
(This article belongs to the Special Issue Porous Carbons for CO2 Adsorption and Capture)

Abstract

:
The CO2 capture from flue gas using biomass-derived porous carbons presents an environmentally friendly and sustainable strategy for mitigating carbon emissions. However, the conventional fabrication of porous carbons often relies on highly corrosive activating agents like KOH and ZnCl2, posing environmental and safety concerns. To address this challenge, in the present work sodium metaborate tetrahydrate (NaBO2·4H2O) has been utilized as an alternative, eco-friendly activating agent for the first time. Moreover, a water chestnut shell (WCS) is used as a sustainable precursor for boron-doped porous carbons with varied microporosity and boron concentration. It was found out that pyrolysis temperature significantly determines the textural features, elemental composition, and CO2 adsorption capacity. With a narrow micropore volume of 0.27 cm3/g and a boron concentration of 0.79 at.% the representative adsorbent presents the maximum CO2 adsorption (2.51 mmol/g at 25 °C, 1 bar) and a CO2/N2 selectivity of 18 in a 10:90 (v/v) ratio. Last but not least, the as-prepared B-doped carbon adsorbent possesses a remarkable cyclic stability over five cycles, fast kinetics (95% equilibrium in 6.5 min), a modest isosteric heat of adsorption (22–39 kJ/mol), and a dynamic capacity of 0.80 mmol/g under simulated flue gas conditions. This study serves as a valuable reference for the fabrication of B-doped carbons using an environmentally benign activating agent for CO2 adsorption application.

1. Introduction

Driven mostly by industrial activities and the continuous dependence on fossil fuels, the constant increase in atmospheric carbon dioxide (CO2) levels (up to 421 ppm) has sharpened the worldwide need for creative and effective carbon capture and storage (CCS) technologies [1,2]. As a first step, CO2 capture from flue gas, or pre-combustion capture, is a primary step to stop the release of CO2 into the atmosphere. Among the many materials such as porous polymers [3,4,5], porous carbons [6,7,8,9,10], zeolites [11] and metal–organic frameworks (MOFs) [12,13] investigated for this use, porous carbon-based sorbent has shown promise because of the strong physicochemical stability, customizable pore shape, large surface area, and advanced adsorption capacity [14,15,16]. Conventional activating agents such as KOH, H3PO4, and ZnCl2 are widely used in the synthesis of porous carbons due to their strong activation capabilities [17,18]. However, their high corrosiveness, hazardous nature, and potential environmental impact present significant challenges in large-scale applications. The use of these aggressive pore-generated agents can lead to equipment degradation, the generation of toxic waste, and safety risks during handling and disposal. Therefore, the development of alternative, eco-friendly activating agents is crucial for promoting sustainable carbon material synthesis. In this context there are significant efforts shown to produce porous carbon using sustainable chemical activation strategies. For instance, Wang et al. reported that weak, corrosive K2CO3 activation enhances porosity, and the resulting adsorbent achieved a CO2 capture capacity of 4.36 mmol/g with good stability [19]. Also, D-glucose was carbonized in a one-step process at 800 °C using potassium acetate, producing porous carbon spheres with an SSA of 1917 m2/g and a specific pore volume of 0.85 cm3/g, achieving a CO2 capture capacity of 6.62 mmol/g at 0 °C and 1 bar [20]. Moreover, shrimp shells were processed too using sodium thiosulfate as an activating agent, yielding ternary (N, S, O)-doped porous carbon with an a CO2 adsorption capacity of 236.80 mg/g and a CO2/N2 selectivity of 84.3 [21]. In addition, in our previous studies we applied sodium thiosulfate (Na2S2O3) [22], potassium thiosulfate K2S2O3 [23], and potassium persulfate [24] as green and less corrosive chemical activating agents. However, to the best of our knowledge there are not any studies using sodium metaborate tetrahydrate (NaBO2·4H2O) as a green activating agent. Based on the MSDSs of KOH and NaBO2·4H2O in Figure S1 (Supplementary Materials), KOH is highly corrosive, causing severe burns, eye damage, and respiratory irritation. KOH is highly reactive with acids and moisture, while NaBO2·4H2O is more stable, with a lower instability rating. This approach not only minimizes hazardous by-products but also enhances the sustainability of porous carbon production by using NaBO2·4H2O as a pore forming agent.
Recent studies have shown that modifying carbon surfaces with heteroatoms significantly enhances CO2 affinity and selectivity through Lewis’s acid–base interactions [25,26,27,28]. Among the various heteroatoms used, nitrogen and sulfur are the most common due to their ability to introduce active sites and alter the electronic properties of the carbon matrix. This modification strengthens the interaction between CO2 molecules and the carbon surface, improving overall adsorption efficiency [29,30,31,32,33]. However, boron doping remains less explored despite its potential to enhance CO2 capture by interacting with the Lewis acid regions of CO2 molecules [34]. In addition to heteroatom doping, the presence of small micropores (<1 nm) further enhances CO2 adsorption by increasing van der Waals forces. This synergistic effect has sparked growing interest in developing biomass-derived porous carbons that integrate optimal porosity with strategic heteroatom incorporation. Such materials offer a sustainable and eco-friendly solution for addressing rising CO2 emissions.
In this regard, agricultural waste offers a plentiful and often underappreciated supply of high-performance carbon sorbents [35,36,37,38]. One particularly interesting option is the lignocellulosic residue of the water chestnut shell (WCS). Commonly discarded and generated in large quantities in regions where water chestnuts are cultivated—especially in Asia and parts of Europe—WCSs contributes to local waste accumulation and environmental burden. Comprising cellulose, hemicellulose, and lignin, its rich composition offers a naturally porous structure that, with heat processing, may be converted into a carbon-rich scaffold perfect for CO2 collection [39]. Selecting water chestnut shell as a forerunner fits the ideas of the circular economy as it transforms an unused waste into a useful instrument for reducing climate change. Moreover, its natural structural characteristics and chemical composition provide a special chance to design materials with customized porosity and usefulness, thereby differentiating it from traditional carbon sources such as coal or synthetic polymers [40].
To address these challenges, the aim of this work is to develop a sustainable and highly efficient boron-doped porous carbon adsorbent derived from water chestnut shell (WCS), using sodium metaborate tetrahydrate (NaBO2·4H2O) as a novel, green activating agent. This study seeks to elucidate the role of boron doping and narrow microporosity on CO2 capture performance, and to demonstrate the potential of WCS as a viable biomass precursor for scalable carbon material production.
Herein, the WCS has been utilized to synthesis boron-doped porous carbons using a two-stage process: first the pre-carbonization of WSCs and then the activation of biochar with sodium metaborate tetrahydrate (NaBO2·4H2O) at elevated temperatures. The resulting materials demonstrate a remarkable potential for CO2 adsorption. These results highlight how closely narrow microporosity, boron doping, and activation conditions interact to control performance. Beyond its capture efficiency, the material shows robustness over many cycles, fast adsorption kinetics, and modest energy needs for regeneration qualities that improve its practical practicality. By using WCSs, this work not only advances the development of environmentally friendly CO2 sorbents but also clarifies the possibility of agricultural wastes to significantly support carbon management strategies, opening the path for more research on biomass-based solutions in the struggle against climate change.

2. Results and Discussion

2.1. Morphology, Structural Phases, and Surface Chemistry Analysis

The SEM images illustrate the morphological transformation of carbonized water chestnut shell (WSC) before and after activation with NaBO2·4H2O. Figure 1a presents the SEM image of WSC, which exhibits a relatively compact, layered structure with irregularly shaped carbon sheets. After activation, as shown in Figure 1b–e, the morphology undergoes significant changes, leading to a more fragmented structure with increased porosity. The activation process induces the formation of interconnected pores and a rougher surface texture, which enhances the surface area and facilitates gas adsorption. Figure 1f is a TEM image of a representative sample WSCSM-750, providing further insight into the nanoscale structure of the activated carbon. A distinctive wormhole-like microstructure is observed, indicating the formation of an intricate network of pores that can improve gas diffusion and adsorption efficiency. Additionally, the absence of well-defined crystalline domains in the TEM image confirms the amorphous nature of the carbon, suggesting that the activation process has effectively disrupted any ordered graphitic structures. This disordered arrangement contributes to enhanced adsorption properties by introducing more active sites and increasing the material’s overall surface area.
The structural characteristics of WSCSM-T samples were further analyzed using X-ray diffraction (XRD, Philips, Almelo, Holland) carried out on a PHILIPS PW3040/60 powder diffractometer using CuKα radiation (λ = 0.15406 nm) and Raman spectroscopy (Reneishaw, Gloucestershire, United Kingdom), examined with a Reneishaw InVia Raman spectrometer using laser excitation wavelengh at 532 nm. The XRD patterns (Figure 2a) exhibit a broad peak at approximately 23.2° corresponding to the (002) plane of carbon, and another wide peak around 43.6° which is assigned to the (100) plane [41,42]. The diffuse nature of these peaks confirms the amorphous structure of the prepared carbons, indicating a lack of long-range order. This observation aligns with the TEM image of WSCSM-750 (Figure 1f), where the absence of well-defined crystalline domains further supports the presence of an amorphous carbon framework. Raman spectroscopy (Figure 2b) provides further insight into the structural evolution of the activated carbons. The spectra display two characteristic peaks of the D-band at approximately 1358 cm−1, associated with structural defects and disorder, and of the G-band at around 1585 cm−1, corresponding to sp2-bonded graphitic carbon. The intensity ratio ID/IG increases from 0.79 for WSCSM-650 to 0.94 for WSCSM-800, indicating a rise in disorder within the carbon framework as the activation temperature increases [43,44]. This trend suggests that the activation process progressively disrupts graphitic domains, creating more defect sites and enhancing porosity. The XRD, Raman, and TEM findings collectively demonstrate that NaBO2·4H2O activation effectively modifies the carbon structure by introducing porosity and disorder. The wormhole-like microstructure observed in TEM indicates the presence of hierarchical porosity, while the increasing ID/IG ratio reflects a more defect-rich carbon matrix at higher activation temperatures. These structural modifications are expected to improve gas diffusion and adsorption capacity, making the WSCSM-T materials promising candidates for CO2 capture applications.
X-ray photoelectron spectroscopy (XPS) was employed to investigate the surface chemistry and elemental composition of the WSCSM-T adsorbents, providing insights into the effectiveness of boron doping. The B 1s XPS spectrum of B-doped porous carbons (Figure 3) exhibits the following three distinct peaks corresponding to different boron bonding states: BC3 at 191.2 eV, BCO2/BC2O at 192.3 eV, and B–O at 193.1 eV. These peaks confirm the successful incorporation of boron into the carbon framework, primarily in the form of boron–carbon and boron–oxygen functionalities [34,45]. Quantitative XPS analysis (Table 1) reveals a progressive increase in boron content from 0.29 at% in WSCSM-650 to 1.65 at% in WSCSM-800. This trend suggests that higher activation temperatures facilitate enhanced boron incorporation, likely due to increased reactivity between the carbon matrix and the NaBO2·4H2O precursor. Notably, the presence of oxygen-bound boron species (BCO2/BC2O and B–O) indicates that boron is not only embedded within the carbon structure but also interacts with surface oxygen groups, potentially influencing the material’s electronic properties and CO2 adsorption performance. The dual role of NaBO2·4H2O, acting as both a pore-forming agent and a boron source, is evident from these results. Its contribution to porosity enhancement, combined with the introduction of boron functionalities, likely improves the interaction between the carbon surface and CO2 molecules. The increasing boron content at elevated temperatures may also contribute to modifications in the electronic structure of the carbon framework, further optimizing its adsorption capabilities. These findings underscore the significance of controlled boron doping in tuning the physicochemical properties of porous carbons for gas capture applications.

2.2. Textural Properties Analysis

The porous textural properties of WSCSM-T were evaluated using N2 adsorption/desorption measurements at −196 °C. As shown in Figure 4a, the sorption isotherms exhibit a distinct Type I behavior, characterized by a steep rise in adsorption at low relative pressures (P/P0 < 0.01), confirming the predominantly microporous nature of the WSCSM-T samples. The semi-plot of the N2 isotherm for WSCSM-T samples is shown in Figure S2 (Supplementary Materials). The textural parameters, including the specific surface area (SBET), total pore volume (V0), and micropore volume (Vt) are summarized in Table 1. With increasing activation temperature, the porosity development follows a distinct trend. WSCSM-700 exhibits the highest SBET (481 m2/g) and the largest micropore volume (V0 = 0.21 cm3/g), indicating that 700 °C is the optimal activation temperature for maximizing microporosity. However, further increasing the activation temperature to 750 °C and 800 °C results in a decline in both surface area and pore volume. Specifically, WSCSM-800 undergoes a significant reduction in SBET to 228 m2/g and a corresponding decrease in micropore volume (V0 = 0.12 cm3/g). This decrease is attributed to structural degradation and pore collapse at elevated temperatures, a phenomenon commonly observed in highly porous carbon materials subjected to excessive thermal treatment. Narrow micropores (<1 nm) play a crucial role in determining CO2 adsorption performance under low-pressure conditions. To assess the contribution of these ultra-micropores, the narrow micropore volumes (Vn) were estimated using the Dubinin–Radushkevich (D–R) equation based on CO2 adsorption data. As listed in Table 1, Vn values range from 0.24 to 0.27 cm3/g, following a trend similar to that observed for SBET and Vt. The pore size distribution (Figure 4b), derived from the density functional theory (DFT) model, provides further insight into the structural characteristics of the WSCSM-T samples. All samples exhibit a dominant micropore size distribution in the range of 1–2 nm, reinforcing their classification as microporous materials. However, with increasing activation temperature, the development of mesopores becomes more apparent, particularly in WSCSM-750 and WSCSM-800. This transition towards a hierarchical pore structure suggests that excessive activation can lead to partial pore widening and the formation of secondary mesopores, which may influence gas diffusion and adsorption kinetics. Overall, these findings highlight the intricate balance between activation temperature and porosity development, where moderate activation temperatures (around 700 °C) favor micropore formation, whereas higher temperatures (≥750 °C) can lead to structural degradation and mesopore generation. The resulting pore architecture is crucial for optimizing CO2 adsorption performance and diffusion efficiency in porous carbon materials.

2.3. CO2 Adsorption Performance

The CO2 adsorption performance of WSCSM-T samples was evaluated at 25 °C and 0 °C, as shown in Figure 5. The adsorption capacities of the samples vary with changes in textural properties and boron content, highlighting the intricate relationship between porosity and surface chemistry in determining CO2 uptake. Narrow micropores (<1 nm) play a dominant role in CO2 adsorption under low pressure due to their strong interactions with CO2 molecules via physisorption. As Vn increases, the CO2 adsorption capacity rises from 2.32 to 2.51 mmol/g at 25 °C and 1 bar which confirms the crucial role of these micropores in enhancing adsorption performance. At 0.15 bar, the typical partial CO2 pressure in post-combustion flue gas, the maximum CO2 uptake for these B-doped porous carbons was 0.98 and 1.65 mmol/g at 25 and 0 °C and 1 bar, respectively. Notably, WSCSM-700 and WSCSM-750 have the same Vn value, yet WSCSM-750 exhibits higher CO2 uptake, likely due to its greater boron content. Boron doping enhances CO2 adsorption by introducing electron-deficient sites, thereby improving the carbon surface’s affinity for CO2 molecules. However, despite WSCSM-800 containing the highest boron content its adsorption capacity does not surpass that of WSCSM-700 or WSCSM-750. This suggests that while boron doping positively influences CO2 uptake it cannot fully compensate for the loss of narrow micropores, which remain the primary adsorption sites. These findings indicate that CO2 uptake is governed primarily by narrow microporosity, which provides high adsorption potential, and secondarily by boron doping, which enhances surface interactions. Therefore, optimizing both structural (narrow microporosity) and chemical (boron doping) properties is essential for maximizing CO2 capture efficiency.
It is worth mentioning that although the maximum CO2 uptake of WSCSM-T samples (2.51 mmol/g) is lower than that of some porous carbons derived from KOH activation [46,47,48,49] it remains comparable to many classical solid adsorbents, such as porous carbons [22,24,50], COFs [51], porous polymers [3], and MOFs [12], among others.
Furthermore, Figure 6 provides a comprehensive evaluation of the CO2 adsorption behavior and separation performance of WSCSM-T samples. Figure 6a presents the CO2 and N2 adsorption isotherms of WSCSM-750 at 25 °C and 1 bar. The material exhibits significantly higher CO2 uptake compared to N2, indicating a strong affinity toward CO2. This enhanced selectivity can be attributed to the combination of an ultra-microporous structure and heteroatom-doped surface chemistry. CO2 has a larger quadrupole moment (–13.4 × 10⁻40 C·m2) and higher polarizability than N2 (–4.7 × 10⁻40 C·m2), which makes it more responsive to electrostatic interactions and Lewis basic sites introduced by boron doping. In addition, the smaller kinetic diameter of CO2 (3.3 Å vs. 3.64 Å for N2) allows it to access narrow micropores more effectively, further contributing to the observed selectivity. As a result, the ideal adsorption solution theory (IAST) [52] predicts a CO2/N2 (10:90, V/V) selectivity of 18, demonstrating the potential of WSCSM-750 for efficient gas separation applications.
The adsorption kinetics curve (Figure 6b) reveals rapid CO2 adsorption, with 95% of the equilibrium capacity achieved within 6.5 min. This fast uptake suggests that WSCSM-750 possesses a well-developed porous structure and high surface accessibility, facilitating efficient gas diffusion. Such rapid adsorption kinetics make WSCSM-750 a strong candidate for cyclic adsorption–desorption processes, which are crucial for reducing energy consumption in industrial CO2 capture and regeneration applications.
Figure 6c illustrates the isosteric heat of adsorption (Qst) for CO2 on WSCSM-T adsorbents, calculated using the Clausius–Clapeyron equation based on isotherm data at 0 °C and 25 °C. The Qst values range from 33 to 39 kJ/mol at near-zero loading, with an overall range of 22–39 kJ/mol. These values suggest that CO2 adsorption is predominantly governed by physisorption, ensuring good regenerability without significant loss of capacity. The relatively high Qst at low coverage also indicates the presence of energetically favorable adsorption sites, likely influenced by heteroatom doping or specific structural features of WSCSM-750.
The breakthrough curve (Figure 6d) further supports the material’s practical applicability, demonstrating a dynamic CO2 capture capacity of 0.80 mmol/g under realistic flow conditions (25 °C, 10 mL/min flow rate, 10 vol.% CO2, and 1 bar total pressure). The combination of fast kinetics, high selectivity, moderate adsorption heat, and stable dynamic capacity underscores WSCSM-750 as a promising and practical adsorbent for CO2 separation processes. It needs to be stated here that there is a certain amount of water vapor present in practical flue gas, which could deteriorate the CO2 capture capacity of these B-doped carbons due to the competitive adsorption between CO2 and H2O.
The cyclic stability of WSCSM-750 as a CO2 sorbent was evaluated over five successive adsorption–desorption cycles (Figure 7). Before each test the sample was heated at 200 °C for 6 h in a vacuum. The results demonstrate remarkable stability, with no significant decrease in adsorption capacity throughout the cycles. This outstanding uniformity highlights the material’s excellent reusability and durability, making it a promising candidate for long-term CO2 capture applications. The sustained performance over multiple cycles suggests that WSCSM-750 maintains its structural integrity and adsorption sites, preventing pore blockage or degradation. This stability is essential for industrial applications, where repeated adsorption and regeneration cycles are required for cost-effective and energy-efficient CO2 capture.

3. Synthesis and Characterization

The dried water chestnut shell was pulverized into a fine powder and subsequently carbonized in a tube furnace. With a continuous N2 flow of 100 mL/min, approximately 5 g of the precursor was placed in a ceramic boat and heated to 500 °C at a rate of 5 °C/min. The preservation of a temperature of 500 °C for two hours ensured complete carbonization. The resultant carbonized product, referred to as WSC, was collected and stored for future use after cooling to ambient temperature in a nitrogen atmosphere.
Boron-doped porous carbons were synthesized through an activation process using NaBO2·4H2O as the activating agent. The WSC and NaBO2·4H2O were mixed in a 1:1 mass ratio, and activation was carried out at four different temperatures: 650, 700, 750, and 800 °C. The activated porous carbons were labeled as WSCSM-T, where T indicates the activation temperature. Detailed procedures for material preparation, physical characterization, and CO2 analysis are provided in the Supplementary Materials.

4. Conclusions

This study successfully synthesized boron-doped porous carbons from water chestnut shells through carbonization and NaBO22O activation at temperatures ranging from 650 to 800 °C. Among the resultant WSCSM-T materials, WSCSM-750 emerged as the optimal sorbent, exhibiting a high narrow micropore volume (0.27 cm3/g) and moderate boron doping (0.79 at%). These structural and chemical features enabled WSCSM-750 to achieve a CO2 adsorption capacity of 2.51 mmol/g at 25 °C and 1 bar, a CO2/N2 selectivity of 18, fast adsorption kinetics, and a dynamic CO2 capture capacity of 0.80 mmol/g. The activation temperature was identified as a crucial factor in balancing narrow microporosity and boron incorporation, which together govern CO2 adsorption performance. These findings underscore the dual role of NaBO2∙4H2O as both an activating and doping agent and highlight the potential of water chestnut shells as a sustainable precursor for high-performance CO2 sorbents. Furthermore, WSCSM-750 demonstrated excellent cyclic stability and moderate Qst values, indicating a well-balanced adsorption strength that ensures both efficient CO2 capture and easy regenerability.
Future research could explore co-doping with additional heteroatoms or fine-tuning activation conditions to further enhance CO2 selectivity and capacity. This study paves the way for scalable, biomass-derived solutions to mitigate CO2 emissions, contributing to the development of sustainable carbon capture technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules30122564/s1, Scheme S1. Schematic diagram of the fixed-bed reactor system, Figure S1: Material safety data sheets (MSDSs) of KOH and NaBO2·4H2O. Figure S2: Semi-plot of the N2 isotherm for WSCSM-T samples.

Author Contributions

Conceptualization, J.W. and X.H.; formal analysis, J.W. and Y.W.; investigation, J.W., Y.W. and X.L.; resources, Q.X., L.W. and X.H.; data curation, Y.L., M.K.A., S.G.C. and M.D.; writing—original draft preparation, J.W. and Q.X.; writing—review and editing, X.H. and Y.L. supervision, X.H. and L.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Zhejiang Provincial Natural Science Foundation (Grant No. LY21B070005), the Science and Technology Innovation Capacity Building Project of the Beijing Academy of Agricultural and Forestry Sciences (Grant No. KJCX20240305), and the Research Support Program for Central Labs at King Khalid University (Project No. CL/CO/B/6).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hanifa, M.; Agarwal, R.; Sharma, U.; Thapliyal, P.C.; Singh, L.P. A review on CO2 capture and sequestration in the construction industry: Emerging approaches and commercialised technologies. J. CO2 Util. 2023, 67, 102292. [Google Scholar] [CrossRef]
  2. Xu, W.-L.; Chen, H.-J.; Wang, Y.-C.; Liu, S.; Wan, X.-Y.; Peng, H.-L.; Huang, K. Chitin-derived fibrous carbon microspheres as support of polyamine for remarkable CO2 capture. Green Chem. Eng. 2022, 3, 267–279. [Google Scholar] [CrossRef]
  3. Sun, L.-B.; Kang, Y.-H.; Shi, Y.-Q.; Jiang, Y.; Liu, X.-Q. Highly Selective Capture of the Greenhouse Gas CO2 in Polymers. ACS Sustain. Chem. Eng. 2015, 3, 3077–3085. [Google Scholar] [CrossRef]
  4. Shao, L.; Li, Y.; Huang, J.; Liu, Y.-N. Synthesis of Triazine-Based Porous Organic Polymers Derived N-Enriched Porous Carbons for CO2 Capture. Ind. Eng. Chem. Res. 2018, 57, 2856–2865. [Google Scholar] [CrossRef]
  5. Sang, Y.F.; Cao, Y.W.; Wang, L.Z.; Yan, W.; Chen, T.W.; Huang, J.H.; Liu, Y.N. N-rich porous organic polymers based on Schiff base reaction for CO2 capture and mercury(II) adsorption. J. Colloid Interface Sci. 2021, 587, 121–130. [Google Scholar] [CrossRef]
  6. Guo, Y.F.; Tan, C.; Sun, J.; Li, W.L.; Zhang, J.B.; Zhao, C.W. Porous activated carbons derived from waste sugarcane bagasse for CO2 adsorption. Chem. Eng. J. 2020, 381, 122736. [Google Scholar] [CrossRef]
  7. Shi, J.; Cui, H.; Xu, J.; Yan, N.; Liu, Y. Design and fabrication of hierarchically porous carbon frameworks with Fe2O3 cubes as hard template for CO2 adsorption. Chem. Eng. J. 2020, 389, 124459. [Google Scholar] [CrossRef]
  8. Shi, S.; Liu, Y. Nitrogen-doped activated carbons derived from microalgae pyrolysis by-products by microwave/KOH activation for CO2 adsorption. Fuel 2021, 306, 121762. [Google Scholar] [CrossRef]
  9. Rehman, A.; Nazir, G.; Rhee, K.Y.; Park, S.-J. A rational design of cellulose-based heteroatom-doped porous carbons: Promising contenders for CO2 adsorption and separation. Chem. Eng. J. 2021, 420, 130421. [Google Scholar] [CrossRef]
  10. Zhi, Y.; Yin, Y.; Ciren, Q.; Xiao, Q.; Zhao, L.; Demir, M.; Simsek, U.B.; Hu, X.; Wang, L. One-step synthesis of K3PO4-activated phosphorus-enriched carbons for enhanced carbon capture. J. Environ. Chem. Eng. 2025, 13, 116694. [Google Scholar] [CrossRef]
  11. Chatterjee, S.; Jeevanandham, S.; Mukherjee, M.; Vo, D.-V.N.; Mishra, V. Significance of re-engineered zeolites in climate mitigation–A review for carbon capture and separation. J. Environ. Chem. Eng. 2021, 9, 105957. [Google Scholar] [CrossRef]
  12. Millward, A.R.; Yaghi, O.M. Metal-Organic Frameworks with Exceptionally High Capacity for Storage of Carbon Dioxide at Room Temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef] [PubMed]
  13. Yu, J.; Xie, L.-H.; Li, J.-R.; Ma, Y.; Seminario, J.M.; Balbuena, P.B. CO2 Capture and Separations Using MOFs: Computational and Experimental Studies. Chem. Rev. 2017, 117, 9674–9754. [Google Scholar] [CrossRef] [PubMed]
  14. Xiao, J.; Yuan, X.; Zhang, T.C.; Ouyang, L.; Yuan, S. Nitrogen-doped porous carbon for excellent CO2 capture: A novel method for preparation and performance evaluation. Sep. Purif. Technol. 2022, 298, 121602. [Google Scholar] [CrossRef]
  15. Wang, Y.; Zhang, C.; Tang, M.; Liu, J.; Yuan, J.; Zhao, Y.; Zhang, G. Preparation of N, S co-doped carbon nanotubes composites by coal pyrolysis for the CO2 capture. J. Environ. Chem. Eng. 2024, 12, 114452. [Google Scholar] [CrossRef]
  16. Yuan, X.; Kumar, N.M.; Brigljević, B.; Li, S.; Deng, S.; Byun, M.; Lee, B.; Lin, C.S.K.; Tsang, D.C.W.; Lee, K.B.; et al. Sustainability-inspired upcycling of waste polyethylene terephthalate plastic into porous carbon for CO2 capture. Green Chem. 2022, 24, 1494–1504. [Google Scholar] [CrossRef]
  17. Quesada-Plata, F.; Ruiz-Rosas, R.; Morallón, E.; Cazorla-Amorós, D. Activated Carbons Prepared through H3PO4-Assisted Hydrothermal Carbonisation from Biomass Wastes: Porous Texture and Electrochemical Performance. ChemPlusChem 2016, 81, 1349–1359. [Google Scholar] [CrossRef]
  18. Pimentel, C.H.; Díaz-Fernández, L.; Gómez-Díaz, D.; Freire, M.S.; González-Álvarez, J. Separation of CO2 using biochar and KOH and ZnCl2 activated carbons derived from pine sawdust. J. Environ. Chem. Eng. 2023, 11, 111378. [Google Scholar] [CrossRef]
  19. Wang, L.; Sun, F.; Hao, F.; Qu, Z.; Gao, J.; Liu, M.; Wang, K.; Zhao, G.; Qin, Y. A green trace K2CO3 induced catalytic activation strategy for developing coal-converted activated carbon as advanced candidate for CO2 adsorption and supercapacitors. Chem. Eng. J. 2020, 383, 123205. [Google Scholar] [CrossRef]
  20. Singh, G.; Ismail, I.S.; Bilen, C.; Shanbhag, D.; Sathish, C.I.; Ramadass, K.; Vinu, A. A facile synthesis of activated porous carbon spheres from d-glucose using a non-corrosive activating agent for efficient carbon dioxide capture. Appl. Energy 2019, 255, 113831. [Google Scholar] [CrossRef]
  21. Nazir, G.; Rehman, A.; Park, S.-J. Valorization of shrimp shell biowaste for environmental remediation: Efficient contender for CO2 adsorption and separation. J. Environ. Manag. 2021, 299, 113661. [Google Scholar] [CrossRef] [PubMed]
  22. Xu, Q.; Wang, J.; Feng, J.; Liu, C.; Xiao, Q.; Demir, M.; Simsek, U.B.; Kılıç, M.; Wang, L.; Hu, X. D-glucose-derived S-doped porous carbon: Sustainable and effective CO2 adsorption. Colloids Surf. A Phys. Eng. Asp. 2025, 709, 136054. [Google Scholar] [CrossRef]
  23. Bai, J.; Shao, J.; Yu, Q.; Demir, M.; Nazli Altay, B.; Muhammad Ali, T.; Jiang, Y.; Wang, L.; Hu, X. Sulfur-Doped porous carbon Adsorbent: A promising solution for effective and selective CO2 capture. Chem. Eng. J. 2024, 479, 147667. [Google Scholar] [CrossRef]
  24. Liu, C.; Zhi, Y.; Yu, Q.; Tian, L.; Demir, M.; Colak, S.G.; Farghaly, A.A.; Wang, L.; Hu, X. Sulfur-Enriched Nanoporous Carbon: A Novel Approach to CO2 Adsorption. ACS Appl. Nano Mater. 2024, 7, 5434–5441. [Google Scholar] [CrossRef]
  25. Quan, C.; Zhou, Y.; Wang, J.; Wu, C.; Gao, N. Biomass-based carbon materials for CO2 capture: A review. J. CO2 Util. 2023, 68, 102373. [Google Scholar] [CrossRef]
  26. Khan, M.H.; Akash, N.M.; Akter, S.; Rukh, M.; Nzediegwu, C.; Islam, M.S. A comprehensive review of coconut-based porous materials for wastewater treatment and CO2 capture. J. Environ. Manag. 2023, 338, 117825. [Google Scholar] [CrossRef]
  27. Karimi, M.; Shirzad, M.; Silva, J.A.C.; Rodrigues, A.E. Biomass/Biochar carbon materials for CO2 capture and sequestration by cyclic adsorption processes: A review and prospects for future directions. J. CO2 Util. 2022, 57, 101890. [Google Scholar] [CrossRef]
  28. Zhi, Y.; Shao, J.; Liu, C.; Xiao, Q.; Demir, M.; Al Mesfer, M.K.; Danish, M.; Wang, L.; Hu, X. High-performance CO2 adsorption with P-doped porous carbons from lotus petiole biomass. Sep. Purif. Technol. 2025, 361, 131253. [Google Scholar] [CrossRef]
  29. Cui, H.; Shi, J.; Xu, J.; Yan, N.; Liu, Y. Direct synthesis of N, S co-doped porous carbons using novel organic potassium salts as activators for efficient CO2 adsorption. Fuel 2023, 342, 127824. [Google Scholar] [CrossRef]
  30. Luo, L.; Yang, C.; Liu, F.; Zhao, T. Heteroatom-N,S co-doped porous carbons derived from waste biomass as bifunctional materials for enhanced CO2 adsorption and conversion. Sep. Purif. Technol. 2023, 320, 124090. [Google Scholar] [CrossRef]
  31. Shao, J.; Wang, Y.; Che, M.; Liu, Y.; Jiang, Y.; Xiao, Q.; Demir, M.; Wang, L.; Hu, X. Sustainable CO2 Capture: N,S-Codoped Porous Carbons Derived from Petroleum Coke with High Selectivity and Stability. Molecules 2025, 30, 426. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, C.; Zhao, T.; Pan, H.; Liu, F.; Cao, J.; Lin, Q. Facile preparation of N-doped porous carbon from chitosan and NaNH2 for CO2 adsorption and conversion. Chem. Eng. J. 2022, 432, 134347. [Google Scholar] [CrossRef]
  33. Li, Y.; Wang, Y.G.; Chen, B.Q.; Wang, L.Y.; Yang, J.; Wang, B.B. Nitrogen-doped hierarchically constructed interconnected porous carbon nanofibers derived from polyaniline (PANI) for highly selective CO2 capture and effective methanol adsorption. J. Environ. Chem. Eng. 2022, 10, 108847. [Google Scholar] [CrossRef]
  34. Shen, Z.; Song, Y.; Yin, C.; Luo, X.; Wang, Y.; Li, X. Construction of hierarchically porous 3D graphene-like carbon material by B, N co-doping for enhanced CO2 capture. Microporous Mesoporous Mater. 2021, 322, 111158. [Google Scholar] [CrossRef]
  35. Li, J.; Michalkiewicz, B.; Min, J.; Ma, C.; Chen, X.; Gong, J.; Mijowska, E.; Tang, T. Selective preparation of biomass-derived porous carbon with controllable pore sizes toward highly efficient CO2 capture. Chem. Eng. J. 2019, 360, 250–259. [Google Scholar] [CrossRef]
  36. Xue, R.; Xu, W.; Ma, X.; Guo, Y.; Zeng, Z.; Li, L.; Si, J. Targeted enhancement of ultra-micropore in highly oxygen-doped carbon derived from biomass for efficient CO2 capture: Insights from experimental and molecular simulation studies. Sep. Purif. Technol. 2025, 353, 128472. [Google Scholar] [CrossRef]
  37. Umar, M.; Yusuf, B.O.; Aliyu, M.; Hussain, I.; Alhassan, A.M.; Awad, M.M.; Taialla, O.A.; Ali, B.; Alhooshani, K.R.; Ganiyu, S.A. Advancing frontiers in CO2 capture: The renaissance of biomass-derived carbon materials. Coord. Chem. Rev. 2025, 526, 216380. [Google Scholar] [CrossRef]
  38. Serafin, J.; Ouzzine, M.; Cruz, O.F.; Sreńscek-Nazzal, J.; Campello Gómez, I.; Azar, F.-Z.; Rey Mafull, C.A.; Hotza, D.; Rambo, C.R. Conversion of fruit waste-derived biomass to highly microporous activated carbon for enhanced CO2 capture. Waste Manag. 2021, 136, 273–282. [Google Scholar] [CrossRef]
  39. Lee, J.; Kim, S.; Lee, K.H.; Lee, S.K.; Chun, Y.; Kim, S.W.; Park, C.; Yoo, H.Y. Improvement of bioethanol production from waste chestnut shells via evaluation of mass balance-based pretreatment and glucose recovery process. Environ. Technol. Innov. 2022, 28, 102955. [Google Scholar] [CrossRef]
  40. Dziejarski, B.; Serafin, J.; Andersson, K.; Krzyżyńska, R. CO2 capture materials: A review of current trends and future challenges. Mater. Today Sustain. 2023, 24, 100483. [Google Scholar] [CrossRef]
  41. Sirinwaranon, P.; Sricharoenchaikul, V.; Vichaphund, S.; Soongprasit, K.; Rodchom, M.; Wimuktiwan, P.; Atong, D. Synthesis and characterization of the porous activated carbon from end-of-life tire pyrolysis for CO2 sequestration. J. Anal. Appl. Pyrolysis 2023, 174, 106139. [Google Scholar] [CrossRef]
  42. Jang, E.; Choi, S.W.; Lee, K.B. Effect of carbonization temperature on the physical properties and CO2 adsorption behavior of petroleum coke-derived porous carbon. Fuel 2019, 248, 85–92. [Google Scholar] [CrossRef]
  43. Goskula, S.; Siliveri, S.; Gujjula, S.R.; Adepu, A.K.; Chirra, S.; Narayanan, V. Development of activated sustainable porous carbon adsorbents from Karanja shell biomass and their CO2 adsorption. Biomass Convers. Biorefinery 2024, 14, 32413–32425. [Google Scholar] [CrossRef]
  44. Nazir, G.; Rehman, A.; Ikram, M.; Aslam, M.; Zhang, T.C.; Khalid, A.; Aftab, S.; Hussain, S.; Almukhlifi, H.A.; Abdel Hafez, A.A.; et al. Facile synthesis of inherently nitrogen-doped PAN-derived microporous carbons activated with NaNH2 for selective CO2 adsorption and separation. Chem. Eng. J. 2024, 497, 154704. [Google Scholar] [CrossRef]
  45. Li, J.; Shi, C.; Bao, A.; Jia, J. Development of Boron-Doped Mesoporous Carbon Materials for Use in CO2 Capture and Electrochemical Generation of H2O2. ACS Omega 2021, 6, 8438–8446. [Google Scholar] [CrossRef]
  46. Dong, J.; Wang, R.; Wang, X.; Tan, S.; Zhao, Z.; Yin, Q.; Lu, X. Preparation of lignin-based porous carbon through thermochemical activation and nitrogen doping for CO2 selective adsorption. Carbon 2024, 229, 119530. [Google Scholar] [CrossRef]
  47. Shi, J.; Xu, J.; Cui, H.; Yan, N.; Yan, R.; Weng, Y. NaCl template synthesis of N-doped porous carbon from magnesium gluconate for efficient CO2 adsorption. Sep. Purif. Technol. 2025, 355, 129756. [Google Scholar] [CrossRef]
  48. Shao, J.; Wang, Y.; Liu, C.; Xiao, Q.; Demir, M.; Al Mesfer, M.K.; Danish, M.; Wang, L.; Hu, X. Advanced microporous carbon adsorbents for selective CO₂ capture: Insights into heteroatom doping and pore structure optimization. J. Anal. Appl. Pyrolysis 2025, 186, 106946. [Google Scholar] [CrossRef]
  49. Shao, J.; Wang, J.; Yu, Q.; Yang, F.; Demir, M.; Altinci, O.C.; Umay, A.; Wang, L.; Hu, X. Unlocking the potential of N-doped porous Carbon: Facile synthesis and superior CO2 adsorption performance. Sep. Purif. Technol. 2024, 333, 125891. [Google Scholar] [CrossRef]
  50. Wu, R.; Bao, A.L. Preparation of cellulose carbon material from cow dung and its CO2 adsorption performance. J. CO2 Util. 2023, 68, 102377. [Google Scholar] [CrossRef]
  51. Furukawa, H.; Yaghi, O.M. Storage of Hydrogen, Methane, and Carbon Dioxide in Highly Porous Covalent Organic Frameworks for Clean Energy Applications. J. Am. Chem. Soc. 2009, 131, 8875–8883. [Google Scholar] [CrossRef] [PubMed]
  52. Myers, A.L.; Prausnitz, J.M. Thermodynamics of mixed-gas adsorption. AIChE J. 1965, 11, 121–127. [Google Scholar] [CrossRef]
Figure 1. SEM images of (a) WSC, (b) WSCSM-650, (c) WSCSM-700, (d) WSCSM-750, and (e) WSCSM-800 and (f) TEM image of WSCPM-750.
Figure 1. SEM images of (a) WSC, (b) WSCSM-650, (c) WSCSM-700, (d) WSCSM-750, and (e) WSCSM-800 and (f) TEM image of WSCPM-750.
Molecules 30 02564 g001
Figure 2. (a) XRD patterns and (b) Raman spectrum graph of B-doped porous carbons derived from WSC.
Figure 2. (a) XRD patterns and (b) Raman spectrum graph of B-doped porous carbons derived from WSC.
Molecules 30 02564 g002
Figure 3. XPS analysis of B 1s spectrum images of (a) WSCSM-650, (b) WSCSM-700, (c) WSCSM-750, and (d) WSCPM-800.
Figure 3. XPS analysis of B 1s spectrum images of (a) WSCSM-650, (b) WSCSM-700, (c) WSCSM-750, and (d) WSCPM-800.
Molecules 30 02564 g003
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution of samples prepared under various conditions. In (a), filled symbols denote the adsorption branches, whereas empty symbols indicate the desorption branches.
Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution of samples prepared under various conditions. In (a), filled symbols denote the adsorption branches, whereas empty symbols indicate the desorption branches.
Molecules 30 02564 g004
Figure 5. CO2 adsorption isotherms for B-doped porous carbons derived from WSC. The isotherms are shown at two distinct temperatures: (a) 25 °C and (b) 0 °C.
Figure 5. CO2 adsorption isotherms for B-doped porous carbons derived from WSC. The isotherms are shown at two distinct temperatures: (a) 25 °C and (b) 0 °C.
Molecules 30 02564 g005
Figure 6. (a) CO2 and N2 adsorption isotherms for WSCSM-750 measured at 25 °C and 1 bar, (b) CO2 adsorption kinetics at 25 °C for WSCSM-750, (c) isosteric heat of CO2 adsorption (Qst) on WSCSM-T adsorbents derived from experimental adsorption isotherms at 0 °C and 25 °C, and (d) breakthrough curves of WSCSM-750 under the following conditions: adsorption temperature of 25 °C, gas flow rate of 10 mL/min, inlet CO2 concentration of 10 vol.%, and gas pressure of 1 bar.
Figure 6. (a) CO2 and N2 adsorption isotherms for WSCSM-750 measured at 25 °C and 1 bar, (b) CO2 adsorption kinetics at 25 °C for WSCSM-750, (c) isosteric heat of CO2 adsorption (Qst) on WSCSM-T adsorbents derived from experimental adsorption isotherms at 0 °C and 25 °C, and (d) breakthrough curves of WSCSM-750 under the following conditions: adsorption temperature of 25 °C, gas flow rate of 10 mL/min, inlet CO2 concentration of 10 vol.%, and gas pressure of 1 bar.
Molecules 30 02564 g006
Figure 7. Cyclic study of CO2 adsorption for WSCSM-750.
Figure 7. Cyclic study of CO2 adsorption for WSCSM-750.
Molecules 30 02564 g007
Table 1. Porous properties, elemental compositions, and CO2 uptakes of precursors and B-doped sorbents derived from different conditions.
Table 1. Porous properties, elemental compositions, and CO2 uptakes of precursors and B-doped sorbents derived from different conditions.
SampleSBET a
(m2/g)
V0 b
(cm3/g)
Vt c
(cm3/g)
Vn d
(cm3/g)
XPS (at. %)CO2 Uptake (mmol/g)
CNBO25 °C0 °C
WSC1160.100.010.191.1884.670.5713.571.562.11
WSCSM-6504640.190.180.242.3087.620.299.782.323.05
WSCSM-7004810.210.190.270.9790.460.408.172.383.27
WSCSM-7504370.190.150.270.9588.430.799.832.513.30
WSCSM-8002280.120.070.260.7886.331.6511.252.332.98
a Surface area was calculated using the BET method at P/P0 = 0.001–0.01. b Total pore volume at P/P0 = 0.99. c Evaluated by the t-plot method. d Pore volume of narrow micropores (<1 nm) obtained from the CO2 adsorption data at 0 °C.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, J.; Wang, Y.; Liu, X.; Xiao, Q.; Demir, M.; Almesfer, M.K.; Colak, S.G.; Wang, L.; Hu, X.; Liu, Y. The Synthesis of B-Doped Porous Carbons via a Sodium Metaborate Tetrahydrate Activating Agent: A Novel Approach for CO2 Adsorption. Molecules 2025, 30, 2564. https://doi.org/10.3390/molecules30122564

AMA Style

Wang J, Wang Y, Liu X, Xiao Q, Demir M, Almesfer MK, Colak SG, Wang L, Hu X, Liu Y. The Synthesis of B-Doped Porous Carbons via a Sodium Metaborate Tetrahydrate Activating Agent: A Novel Approach for CO2 Adsorption. Molecules. 2025; 30(12):2564. https://doi.org/10.3390/molecules30122564

Chicago/Turabian Style

Wang, Junting, Yingyi Wang, Xiaohan Liu, Qiang Xiao, Muslum Demir, Mohammed K. Almesfer, Suleyman Gokhan Colak, Linlin Wang, Xin Hu, and Ya Liu. 2025. "The Synthesis of B-Doped Porous Carbons via a Sodium Metaborate Tetrahydrate Activating Agent: A Novel Approach for CO2 Adsorption" Molecules 30, no. 12: 2564. https://doi.org/10.3390/molecules30122564

APA Style

Wang, J., Wang, Y., Liu, X., Xiao, Q., Demir, M., Almesfer, M. K., Colak, S. G., Wang, L., Hu, X., & Liu, Y. (2025). The Synthesis of B-Doped Porous Carbons via a Sodium Metaborate Tetrahydrate Activating Agent: A Novel Approach for CO2 Adsorption. Molecules, 30(12), 2564. https://doi.org/10.3390/molecules30122564

Article Metrics

Back to TopTop