Next Article in Journal
Formation of Improved Metallurgical Properties and Carbon Structure of Coke by Optimizing the Composition of Petrographically Heterogeneous Interbasin Coal Batches
Previous Article in Journal
Atomic-Scale Mechanisms of Catalytic Recombination and Ablation in Knitted Graphene Under Hyperthermal Atomic Oxygen Exposure
Previous Article in Special Issue
Biochar in Agriculture: A Review on Sources, Production, and Composites Related to Soil Fertility, Crop Productivity, and Environmental Sustainability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

NaOH-Modified Activated Carbon Materials for Hydrogen Sulfide Removal

by
Meriem Abid
,
Manuel Martínez-Escandell
and
Joaquín Silvestre-Albero
*
Laboratorio de Materiales Avanzados, Departamento de Química Inorgánica, Instituto Universitario de Materiales, Universidad de Alicante, 03690 San Vicente del Raspeig, Spain
*
Author to whom correspondence should be addressed.
Submission received: 7 May 2025 / Revised: 8 August 2025 / Accepted: 31 August 2025 / Published: 3 September 2025
(This article belongs to the Special Issue Carbons for Health and Environmental Protection (2nd Edition))

Abstract

A high-surface-area activated carbon material (RG) is used as a platform to create highly concentrated NaOH composites. These materials are tested for the removal of H2S under industrially relevant conditions (800 ppm H2S in CO2-, H2O- and O2-containing streams). The experimental results show that the breakthrough performance highly depends on the amount of NaOH incorporated and the experimental conditions used (e.g., relative humidity). The most promising material (RG-NaOH-30) reaches a saturation uptake of up to 800 mgH2S/g at 25 °C and atmospheric pressure. This value is among the most promising results reported in the literature for H2S removal, and it is well above traditional commercial samples. Breakthrough column tests confirm the promoting role of humidity in the reaction mechanism. Analysis of the adsorbents after H2S confirms the formation of well-defined sulfur (Sn) microcrystals as the main reaction product.

Graphical Abstract

1. Introduction

Activated carbon is considered a widely versatile adsorbent platform and constitutes the primary class of physical adsorbent used in industry, after activated alumina, with a total market of USD 4.4 billion in 2023. Activated carbon materials are widely used in gas- and liquid-phase adsorption processes (e.g., water treatment, food and beverages, pharmaceuticals, etc.). The widespread use of activated carbon materials is based on their widely developed porous structure and tailored surface chemistry [1]. While carbon materials exhibit an excellent performance for high-boiling-point molecules (e.g., volatile organic compounds), these materials fail for low-boiling-point probes (e.g., toxic or odorous gases) due to the small adsorption potential and low packing density in narrow pores, as well as the absence of condensation processes (supercritical conditions) [2,3,4]. These limitations can be overcome either by using high pressure (e.g., methane storage) or through the incorporation of an impregnant able to selectively trap the target compound (e.g., the removal of toxic gases) [1,5,6]. One of these gases is hydrogen sulfide (H2S). Hydrogen sulfide is a component in hydrocarbon sources (e.g., natural gas, biogas, syngas, etc.), and must be selectively removed to mitigate associated side effects such as corrosion in industrial facilities, poison for catalysts, environmental pollution, and health effects [7]. Among technological approaches, adsorption using impregnated carbon materials is one of the most promising alternatives due to the low cost and high efficiency of the process (unfortunately, unmodified carbon materials exhibit a poor adsorption performance) [8,9]. The development of high-surface-area carbon materials as a support for the active phase (e.g., either metal oxides or caustic impregnants) is of paramount importance to provide high dispersion, proper diffusion, and consequently, an optimized performance [10,11]. Impregnation with sodium hydroxide, potassium hydroxide, sodium carbonate, potassium carbonate, potassium iodide, and potassium iodate has been widely described in the literature for H2S removal [11,12,13,14,15]. These studies have anticipated the crucial role of the alkali impregnant in the removal process, although it is associated with severe blocking effects after deposition and/or after the H2S adsorption process, with the associated formation of bulky carbonates or sulfates. Another limitation arises when comparing activated carbons from different origins due to the concomitant effect of a different porosity (promoting H2S physisorption), and the associated differences in the characteristics of the alkaline species deposited (promoting H2S chemisorption/reaction) [16]. The presence of these variables simultaneously prevents proper evaluation of the real role exerted by the alkali impregnant, triggering important questions regarding the role of alkali content, effect of humidity, and effect of the intrinsic porosity of the carbon material, among others. Furthermore, a proper evaluation of the exhausted/used catalysts is needed to provide more insight about the reaction mechanism for NaOH-impregnated carbons (e.g., formation of elemental sulfur, sulfur dioxide, sulfates, or even sulfuric acid).
With these premises in mind, the objective of this study is the development of novel H2S adsorbents using sodium hydroxide (NaOH) as an impregnant and a common activated carbon material as a support. The activated carbon selected is characterized by a highly developed porous structure (large BET surface area) and a perfectly tailored micro-/mesoporous network to minimize blocking effects. This material will be used as a support for different NaOH loadings to be applied under dry and moist conditions under industrially relevant H2S streams.

2. Materials and Methods

High-surface-area RGC-30 activated carbon (MEAD-Westvaco Corp., Richmond, VA, USA) was selected as the support material for this study. Sodium hydroxide (NaOH) was purchased from Sigma-Aldrich (reagent grade, ≥98%).

2.1. Preparation of Adsorbents

Different adsorbents were prepared using a wet impregnation approach. Before the impregnation step, the activated carbon sample was thermally treated at 70 °C overnight in order to remove any residual adsorbed species. In total, 1 g of activated carbon RGC-30 (particle size 0.5 mm) was added to a 10 mL deionized water solution containing corresponding amounts of NaOH (10 wt.%, 15 wt.%, 20 wt.%, 30 wt.%, and 40 wt.%; theoretical weight % per gram of carbon). The main goal of using highly concentrated NaOH solutions was to reach the maximum loading capacity and, indirectly, the maximum H2S removal performance. After 1 h under stirring, the samples were filtered and dried at 70 °C for 24 h. The real NaOH content was calculated from the final weight of the pre-impregnated activated carbon. The samples were labeled RG-NaOH-10, RG-NaOH-15, RG-NaOH-20, RG-NaOH-30, and RG-NaOH-40, where the number represents the nominal NaOH content. The blank sample without NaOH was labeled RG.

2.2. Samples Characterization

The textural characteristics of the activated carbon materials were evaluated by nitrogen adsorption isotherms at cryogenic temperatures. Gas physisorption analysis was performed in homemade fully automated manometric equipment designed and constructed by the LMA research group. Prior to the measurements, the samples were outgassed at 250 °C for 4 h. The apparent surface area was calculated using the BET equation.
Raman spectra were obtained in a JascoNRS-5100 equipment (Tokyo, Japan) using a 532 nm laser and 600 lines per mm slit between 0 and 4000 cm−1. Surface morphology was evaluated using FE-SEM, and mapping analysis was performed using energy-dispersive X-ray spectroscopy (EDX). XRD patterns were measured in a Bruker diffractometer (Billerica, MA, USA), model D8-Advance, provided with a copper anode and a Goebel mirror. The samples were scanned between 5° and 80° with a pre-set of 3 s, an angular speed of 1°/min, and steps of 0.05. XPS analyses were performed with a K-ALPHA ThermoScientific instrument (Waltham, MA, USA) at a pressure of 5 × 10−7 N·m−2. The spectra were obtained using Al K radiation (1486.6 eV) with a twin crystal monochromator.
pH measurements were carried out by mixing the sample (0.4 g) with ultrapure water (20 mL) and stirring the suspension for 24 h to reach equilibrium. The pH of the solution was measured at 25 °C. Thermogravimetric analyses were carried out in Mettler Toledo TG-DSC2 equipment (Greifensee, Switzerland) using 20 mg of sample in an alumina crucible and a thermal treatment up to 1000 °C in nitrogen, accordingly (100 mL/min; heating rate 10 °C/min).

2.3. H2S Removal Process

The H2S removal performance was tested at room temperature (25 °C) using a fixed-bed tubular reactor (i.d., 1 cm) and a total flow of 80 mL/min. This flow was divided in two streams, (i) 40 mL/min of a N2/O2 (5%O2) mixture (cylinder I) and (ii) 40 mL/min of a H2S (1600 ppm)/N2 (50%)/CO2 (50%) (cylinder II), at ambient temperature and atmospheric pressure. Breakthrough experiments were performed using 1 g of sample and an electrochemical detector (Dräger-polytron) for continuously recording the H2S concentration. Breakthrough column experiments were performed in dry and wet conditions (40–70% relative humidity range) using a thermostatic bath fixed at 10 –17 °C (depending on the desired relative humidity), while bubbling the effluent from cylinder I. All experiments were repeated twice to check the reproducibility in the breakthrough tests. Used samples were characterized as described in Section 2.2 and labeled “-af”.

3. Results and Discussion

The textural properties of the original and NaOH-loaded samples are evaluated using nitrogen adsorption measurements at cryogenic temperatures. Figure 1a confirms the micro-/mesoporous nature of the original RG activated carbon with a specific BET surface area of 1425 m2/g. Impregnation with NaOH gives rise to a decrease in the nitrogen adsorption capacity, mainly in the microporous region. As shown in Figure 1a and Table S1, the amount of nitrogen adsorbed decreases with the amount of NaOH incorporated, down to a BET surface area of 1140 m2/g for sample RG-NaOH-40 (~20% decrease). This decrease in textural properties after NaOH incorporation does not necessarily reflect the presence of blocking effects by NaOH, but can also be attributed to the additional weight of the impregnated samples. Indeed, the normalization of the nitrogen adsorption isotherms per unit mass of activated carbon (Figure S1) confirms the absence of severe blocking effects by NaOH, i.e., the deposition of NaOH on the surface of the high-surface-area RG activated carbon creates a thin alkaline film with basic properties, with minimal porosity loss. The formation of an extended interfacial layer while preserving most of the excellent textural properties (micro- and mesoporous structure) in a high-surface-area porous system is an ideal scenario to obtain optimized H2S adsorbents. Pore size distribution estimated from the nitrogen adsorption data (Figure 1b) confirms the combined presence of narrow and wide micropores in the evaluated carbons, together with some significant contributions of mesopores around 3–5 nm, with small differences between the samples. The total amount of NaOH incorporated is roughly estimated from the weight increase after impregnation. The real NaOH content ranges from 3.7 wt.% for sample RG-NaOH-10 up to 13.4 wt.% for sample RG-NaOH-40 (Table S1). These results anticipate that the total amount of NaOH that can be incorporated in a high-surface-area carbon material, such as RG, is limited, even at extremely high solution concentrations (nominal value of 40 wt.% vs. real value of 13.4 wt.% for sample RG-NaOH-40).
The Raman spectra of the modified carbon materials shown in Figure 2a are dominated by two intense bands at ~1350 cm−1 and 1600 cm−1, which are attributed to D and G contributions, respectively. More specifically, the D band corresponds to the A1g breathing mode symmetry, which is associated with the carbon atoms having a disordered sp3 configuration, while the G band is associated with the E2g mode symmetry, corresponding to the sp2 carbon system, and results from the in-plane bond stretching of the C-C bond in graphitic materials [17]. The ID/IG ratio (an indirect measure of structural disorder) is rather similar among the samples, thus suggesting that NaOH incorporation does not modify the carbon matrix. The inset reflects the new Raman shifts developed after NaOH incorporation, attributed to O-H vibrational modes from NaOH or H2O in the range of 2000–3500 cm−1. XRD patterns for the pristine activated carbon (Figure 2b) reflect characteristic broad diffraction peaks at 23° and 43°, attributed to the (022) and (100) planes from the graphitic microdomains in the amorphous activated carbon structure. The incorporation of NaOH does not produce any modification in the XRD pattern, except for in samples RG-NaOH-30 and RG-NaOH-40, with new diffraction peaks appearing in the 2-theta range of 30–50°. These new peaks do not correspond to NaOH, but rather to Na2CO3 microcrystals formed on the carbon surface (main peaks at 30.2°, 34°, 35.2°, and 38°) [18,19]. These results anticipate that the thermal treatment at 70 °C performed after impregnation with NaOH promoted the carbonation reaction with ambient CO2 to Na2CO3 [20]. However, the presence of amorphous NaOH cannot be ruled out.
For a better understanding of the morphology and microstructure of the NaOH-modified carbon materials, the samples are evaluated using FE-SEM. Representative images (Figure 3) show the granular shape of the pristine RG activated carbon and the formation of white needles in the modified samples due to the crystallization of Na2CO3, in close agreement with the XRD patterns. The good dispersion of the Na2CO3 nano-crystallites on the carbon structure and the absence of appreciable aggregation (or blocking) constitute a priori an excellent platform for the design of efficient H2S adsorbents. Mapping analysis (Figure S2) confirms the excellent dispersion of sodium atoms in the evaluated surfaces.
In general, hydrogen sulfide removal using activated carbon materials involves several steps, i.e., the adsorption of H2S on the carbon surface, the dissolution and dissociation of H2S on the surface water film (under humid conditions), facilitating the formation of HS and H+, the oxidation of HS with adsorbed oxygen, and the formation of elemental sulfur or sulfur dioxide (depending on the surface characteristics). The additional oxidation of SO2 to H2SO4 can take place in the presence of water and metal impurities that promote catalytic oxidation [14]. Therefore, the removal mechanism combining physical adsorption and a subsequent oxidation step can be summarized as follows:
H 2 S g H 2 S   ( a d s )
H 2 S a d s H 2 S   ( a d s l i q )
H 2 S a d s l i q H S   ( a d s ) + H +
O 2 g O 2   ( a d s )
O 2 a d s O 2   ( a d s l i q )
O 2 a d s l i q 2 O *   ( a d s )
H S a d s + O * a d s S   ( a d s ) + O H
H S a d s + 3   O * a d s S O 2   ( a d s ) + O H
S O 2 a d s + O * a d s + H 2 O H 2 S O 4   ( a d s )
H + + O H H 2 O
An important factor that plays a significant role in the desulfurization process is the surface pH of the carbon before and after impregnation. A low pH of the carbon surface is expected to suppress H2S dissociation and the creation of HS ions so that only physical adsorption can occur. Under these conditions, sulfur is easily oxidized to SO2 and converted to H2SO4 in small pores [14]. On the contrary, a basic pH promotes these dissociation processes to produce HS or S2−, with HS being oxidized to polymeric sulfur with a chain or ring-like morphology (Sn). The pH values obtained in this study are reported in Table 1. While the pristine carbon has a neutral pH, loading NaOH gives rise to a significant increase in the pH up to values of around 10.3–10.6, independently of the amount of NaOH incorporated.
The pKa constants for H2S are 7.2 and 13.9 for the first and second dissociation, respectively. Based on the pH values obtained, a high concentration of HS ions is expected under the experimental conditions tested in this study, including the possibility to make polysulfides.
The thermal behavior of the NaOH-loaded samples is evaluated using thermogravimetric analysis. Figure 4 shows the TG profiles for the different samples in the temperature range of 25–1000 °C under a nitrogen atmosphere. The TG profiles exhibit an initial weight loss at low temperatures (around 80–100 °C) due to the removal of moisture. Above 100 °C, the TG profile remains relatively stable up to 700–800 °C, where a significant weight loss is observed (up to 1000 °C). The magnitude of this weight loss at high temperatures scales up with the amount of NaOH incorporated (5 wt.% for RG; 12 wt.% for RG-NaOH-10; 13 wt.% for RG-NaOH-15; 15 wt.% for RG-NaOH-20; 18 wt.% for RG-NaOH-30; and 20 wt.% for RG-NaOH-40), most probably associated with the high-temperature decomposition of the alkali carbonate (Na2CO3) to CO2 and Na2O. Na2CO3 decomposition is traditionally associated with a large CO2 evolution starting at 780 °C and reaching the maximum at 850 °C, in perfect agreement with our TG measurements [21].
The performance of the NaOH-loaded samples in the removal of H2S was tested under dry and wet conditions. To this end, breakthrough column experiments were performed using a simulated industrial stream containing 800 ppm H2S, 2.5% O2, and 50% CO2. Figure 5a reports the breakthrough column performances for all the NaOH-loaded samples under humid conditions (60% relative humidity) at 25 °C. The performance of the pristine activated carbon is very limited under these experimental conditions, with a total uptake as low as 2 mg/g. The result achieved with sample RG anticipates that the presence of a highly developed porous structure and a large pore volume in the microporous range is not sufficient to reach a high H2S adsorption capacity [14,22]. In other words, the role of micropores and mesopores in the H2S removal process, if any, must be very limited. On the contrary, samples modified with NaOH exhibit a large improvement in removal performance. For all samples, the breakthrough profile at saturation exhibits a sharp increase, characteristic of acid–base reaction processes, i.e., once all basic sites are saturated, H2S breaks the column sharply. As expected, the breakthrough time and, indirectly, the total removal capacity scale-up with the amount of NaOH incorporated up to an optimum value for sample RG-NaOH-30. Larger NaOH contents become detrimental for the total removal efficiency, most probably due to the growth of large Na2CO3 nanocrystals (less optimized dispersion). The total removal capacity measured up to the breakthrough point is reported in Figure 5b. Highly dispersed Na2CO3 nanocrystals in RG activated carbon give rise to a large improvement in the H2S removal capacity, e.g., a more than one hundred times (100×) improvement between unmodified RG and sample RG-NaOH-10. Several repetitions are performed to evaluate the reproducibility of the obtained results, with the standard deviation among measurements being very small for all samples evaluated. This is somehow expected, because NaOH impregnation must be relatively homogeneously dispersed, giving rise to an optimized performance. At this point, it is important to highlight that the total removal capacity for samples RG-NaOH-30 and RG-NaOH-40 is highly above the threshold value of 600 mg/g. Sample RG-NaOH-30 even reaches values close to 800 mg/g. To our knowledge, this is one of the best adsorption values reported in the literature for H2S removal using either catalytic or acid–base adsorbents [23]. The excellent performance achieved with these samples must be attributed to the presence of highly dispersed NaOH/Na2CO3 nanocrystals on the surface of a highly activated carbon material, such as RG. Previous studies described in the literature using Fe@carbon-rich nanoparticles reported exceptional values slightly above 600 mg/g under similar experimental conditions [23]. The value obtained with sample RG-NaOH-30 highly surpasses the performance of catalytic Fe@carbon nanoparticles and constitutes a four-fold increase compared to commercial H2S adsorbents (chemically impregnated adsorbents) such as ADDSORBTM VA3 and ADDSORBTM VA6, with total values of 177 and 204 mg/g, respectively, under the same experimental conditions (in agreement with the values reported in the technical sheets) [23]. The obtained values constitute a four-fold increase compared to the theoretical predictions for MOFs (e.g., MIL-101) in the H2S process, and are well above the performance of metal oxides and zeolites [24,25]. Overall, these results confirm that the newly designed NaOH-impregnated carbons exhibit exceptional behavior for H2S removal. Taking into account the stoichiometry of the acid–base reaction process, the obtained results are twelve times above the stoichiometric value for the real amount of NaOH incorporated. This observation reflects the autocatalytic behavior of the reaction products, as described in the literature [14].
One of the critical parameters in the removal of H2S is the presence or absence of humidity and the potential formation of a water film on the carbon surface. To evaluate the effect of relative humidity on the H2S removal performance, sample RG-NaOH-30 is tested under different moisture conditions. As can be appreciated in Figure 6, moisture content is critical for the removal process, with the total amount adsorbed ranging from 307 mg/g at 40% RH up to 800 mg/g at 60% relative humidity. Unexpectedly, a larger humidity content becomes detrimental for H2S adsorption performance, with a significant decrease (~30% reduction). In any case, the breakthrough tests exhibit a similar profile shape, independently of the moisture concentration.
To obtain some insights into the reaction mechanism, used samples (after H2S) were evaluated using FE-SEM (samples labeled RG-NaOH-XX-af). Figure 7 displays the morphology of the adsorbents after being exposed to saturation with hydrogen sulfite. Nicely, the FE-SEM images show the formation of large micron-size crystals, more clearly visible in the samples with a high NaOH content. The presence of perfectly defined microcrystals must be related to the reaction between NaOH and H2S, most probably associated with the formation of sodium sulfate crystals (Na2SO4) and/or Sn crystals. The formation of sodium sulfate could be due to the interaction between SO2 and NaOH or Na2CO3, through the following reactions:
N a 2 C O 3 + S O 2 + 1 2   O 2 N a 2 S O 4 + C O 2
2   N a O H + S O 2 N a 2 S O 3 + H 2 O
N a 2 S O 3 + 1 2   O 2 N a 2 S O 4
The formation of Sn will be associated with the following reactions:
2   H 2 S + S O 2 3 S + 2   H 2 O
N a O H + H 2 S N a H S + H 2 O
N a 2 C O 3 + 2   H 2 S 2   N a H S + C O 2 + H 2 O
N a H S + 1 2   O 2 S + N a O H
The predominant gemstone-like shape of the observed crystals suggests the presence of sulfur (Sn) crystals, rather than Na2SO4 (more needle-like crystals will be formed). Furthermore, the associated formation of NaOH upon NaHS oxidation will explain the large efficiency of the designed sorbents, well above the stoichiometric values.
Post mortem sorbents were also evaluated using X-ray diffraction. The XRD patterns described in Figure S3 keep the broad bands at 23° and 45° due to the graphitic structure. Contrary to the fresh samples, the XRD peaks corresponding to Na2CO3 vanish after H2S exposure and new bands emerge at 22.5°, 23.6°, 24.7°, 31°, and 32.7°. Although these bands could be attributed to Na2SO4, they better fit with the fingerprint of S8 crystals formed on the material surface [26,27]. These peaks are more clearly appreciated for sample RG-NaOH-40-af.
Used adsorbents were also evaluated using thermogravimetric analyses (Figure S4). The TG profiles of the NaOH-loaded samples are characterized by a prominent peak at ca. 360–400 °C and some minor contributions at 700–800 °C. The main contribution at low temperatures is not present in the original samples and, consequently, must be derived from the H2S removal process. Furthermore, the main contribution presents a small shoulder at lower temperatures (~265 °C), only visible in the samples loaded with NaOH. Previous studies described in the literature have anticipated that the desorption of oxidized sulfur species (SO2, H2SO4, etc.) takes place at lower temperatures (<300 °C), while higher temperatures (400–500 °C) correspond to the desorption of elemental sulfur [14,22]. The increased magnitude of the desorption peak at 360–400 °C with the amount of NaOH incorporated is in close agreement with the improved adsorption performance for samples with high caustic loadings and the higher S8 formation yield. Furthermore, the shoulder at 265 °C due to oxidized sulfur is rather small and similar among all the samples evaluated, thus confirming that the formation of SO2 and H2SO4 must be quite restricted under the experimental conditions tested. High-temperature peaks (>700 °C) must be associated with other sulfur species, most probably located in small micropores. The pH of the used samples (Table 1) after the H2S breakthrough experiment ranges from 3.5 to 8.7, depending on the NaOH concentration, thus confirming the preferential formation of sulfur. However, the formation of some residual highly acidic sulfur species (e.g., H2SO4) cannot be completely ruled out, in agreement with the TG measurements.
To further clarify the reaction mechanism and identify the nature of the observed crystals, the used adsorbents are evaluated using X-ray photoelectron spectroscopy (XPS). Figure 8 shows representative spectra for the used sample RG-NaOH-30-af in the C1s, O1s, S2p, and Na1s regions. Quantitative analysis of the different materials is reported in Table S2. As expected, the used sorbents exhibit a large sulfur content ranging from 8 at.% for sample RG-NaOH-10 up to 17–18 at.% for samples RG-NaOH-30/-40. A closer look at the XPS spectra in the S2p region (Figure 8) clearly identifies two predominant peaks attributed to sulfur (S8), with binding energies at 164.0 and 165.3 eV, while the contribution of sulfates (above 168 eV) is very small [23,28,29]. These results confirm that the nature of the perfectly shaped microcrystals observed using the FE-SEM images of the used adsorbents corresponds to S8 gemstone-like crystals, and not to Na2SO4.

4. Conclusions

The results presented in this paper show that activated carbon materials with a widely developed porous structure (containing micro- and mesopores) constitute an excellent platform to reach high NaOH loadings. These materials outperform commercial adsorbents for H2S removal under an industrially relevant environment. Breakthrough column tests confirm the excellent performance of these systems, with adsorption values close to 800 mg/g. These results confirm the promoting role of humidity in the reaction mechanism, through the formation of HS species, and elemental sulfur as the main reaction product. Furthermore, breakthrough column tests confirm the beneficial effect of NaOH incorporation on removal performance, with an optimum performance observed for sample RG-NaOH-30. The presence of well-faceted sulfur crystals as reaction products is confirmed by FESEM and XPS. Despite the promising results achieved so far, more research is needed to identify potential regeneration approaches.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/c11030068/s1, Figure S1: Nitrogen adsorption/desorption isotherms at −195 °C normalized per gram of activated carbon (using the real amount of NaOH incorporated); Figure S2: Elemental mapping for C (red) and Na (blue) of the synthesized RG-NaOH-X samples; Figure S3: XRD patterns of the different adsorbents after the H2S adsorption process; Figure S4: Thermogravimetric analysis of the used adsorbents after the hydrogen sulfide breakthrough tests; Table S1: Textural properties of the support and the impregnated samples from the N2 adsorption measurements; Table S2: Chemical composition for the different samples evaluated after H2S experiments deduced from the XPS data.

Author Contributions

Conceptualization, J.S.-A. and M.M.-E.; methodology, J.S.-A.; validation, M.A.; formal analysis, J.S.-A.; investigation, M.A.; data curation, M.A.; writing—original draft preparation, J.S.-A. and M.M.-E.; writing—review and editing, J.S.-A.; supervision, J.S.-A.; project administration, J.S.-A.; funding acquisition, J.S.-A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by MCIN (Project PID2022-142960OB-C21), Conselleria de Innovación, Universidades, Ciencia y Sociedad Digital, Generalitat Valenciana (Project CIPROM/2021/022), and EU MSCA (Project CLEANWATER; Grant Agreement: 101131382).

Data Availability Statement

The data presented in this study area available on request from the corresponding author due to legal reasons.

Acknowledgments

M.A. acknowledges financial support from the Algerian Ministry of Higher Education and Scientific Research.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Marsh, H.; Rodríguez-Reinoso, F. Activated Carbon; Elsevier: Amsterdam, The Netherlands, 2006. [Google Scholar]
  2. Sabzehmeidani, M.M.; Mahnaee, S.; Ghaedi, M.; Heidari, H.; Roy, V.A.L. Carbon Based Materials: A Review of Adsorbents for Inorganic and Organic Compounds. Mater. Adv. 2021, 2, 598–627. [Google Scholar] [CrossRef]
  3. Singh, R.; Wang, L.; Ostrikov, K.; Huang, J. Designing Carbon-Based Porous Materials for Carbon Dioxide Capture. Adv. Mater. Interfaces 2024, 11, 2202290. [Google Scholar] [CrossRef]
  4. Silvestre-Albero, A.; Ramos-Fernandez, J.M.; Martinez-Escandell, M.; Sepúlveda-Escribano, A.; Silvestre-Albero, J.; Rodríguez-Reinoso, F. High Saturation Capacity of Activated Carbons Prepared from Mesophase Pitch in the Removal of Volatile Organic Compounds. Carbon 2010, 48, 548–556. [Google Scholar] [CrossRef]
  5. Rodríguez-Reinoso, F.; Silvestre-Albero, J. Methane Storage on Nanoporous Carbons. In Nanoporous Materials for Gas Storage; Kaneko, K., Rodríguez-Reinoso, F., Eds.; Springer Nature Singapore Pte Ltd.: Singapore, 2019; pp. 209–226. [Google Scholar]
  6. Johnson, R.L.; Kuzub, R.E.; Tak, J.K. Acid-Impregnated Activated Carbon and Methods of Forming and Using the Same. U.S. Patent US9114358B2, 25 August 2015. [Google Scholar]
  7. Nuvolone, D.; Petri, D.; Pepe, P.; Voller, F. Health Effects Associated with Chronic Exposure to Low-Level Hydrogen Sulfide from Geothermoelectric Power Plants: A Residential Cohort Study in the Geothermal Area of Mt. Amiata in Tuscany. Sci. Total Environ. 2019, 659, 973–982. [Google Scholar] [CrossRef]
  8. Khabazipour, M.; Anbia, M. Removal of Hydrogen Sulfide from Gas Streams Using Porous Materials: A Review. Ind. Eng. Chem. Res. 2019, 58, 22133–22164. [Google Scholar] [CrossRef]
  9. Pudi, A.; Rezaei, M.; Signorini, V.; Andersson, M.P.; Baschetti, M.G.; Mansouri, S.S. Hydrogen Sulfide Capture and Removal Technologies: A Comprehensive Review of Recent Developments and Emerging Trends. Sep. Purif. Technol. 2022, 298, 121448. [Google Scholar] [CrossRef]
  10. Cimino, S.; Lisi, L.; de Falco, G.; Montagnaro, F.; Balsamo, M.; Erto, A. Highlighting the Effect of the Support During H2S Adsorption at Low Temperature over Composite Zn-Cu Sorbents. Fuel 2018, 221, 374–379. [Google Scholar] [CrossRef]
  11. Bagreev, A.; Bandosz, T.J. A Role of Sodium Hydroxide in the Process of Hydrogen Sulfide Adsorption/Oxidation on Caustic-Impregnated Activated Carbons. Ind. Eng. Chem. Res. 2002, 41, 672–679. [Google Scholar] [CrossRef]
  12. Sitthikhankaew, R.; Chadwick, D.; Assabumrugrat, S.; Laosiripojana, N. Effect of KI and KOH Impregnations Over Activated Carbon on H2S Adsorption Performance at Low and High Temperatures. Sep. Sci. Technol. 2014, 49, 354–366. [Google Scholar] [CrossRef]
  13. Xiao, Y.; Wang, S.; Wu, D.; Yuan, Q. Experimental and Simulation Study of Hydrogen Sulfide Adsorption on Impregnated Activated Carbon Under Anaerobic Conditions. J. Hazard. Mater. 2008, 153, 1193–1200. [Google Scholar] [CrossRef]
  14. Yan, R.; Chin, T.; Ling Ng, Y.; Duan, H.; Liang, D.T.; Tay, J.H. Influence of Surface Properties on the Mechanism of H2S Removal by Alkaline Activated Carbons. Environ. Sci. Technol. 2004, 38, 316–323. [Google Scholar] [CrossRef]
  15. Sitthikhankaew, R.; Predapitakkun, S.; Kiattikomol, R.; Pumhiran, S.; Assabumrungrat, S.; Laosiripojana, N. Comparative study of hydrogen sulfide adsorption by using alkaline impregnated activated carbons for hot fuel gas purification. Energy Proc. 2011, 9, 15–24. [Google Scholar] [CrossRef]
  16. Yan, R.; Liang, D.T.; Tsen, L.; Tay, J.H. Kinetics and mechanism of H2S adsorption by alkaline activated carbon. Environ. Sci. Technol. 2002, 36, 4460–4466. [Google Scholar] [CrossRef]
  17. Wang, Y.; Alsmeyer, D.C.; McCreery, R.L. Raman Spectroscopy of Carbon Materials: Structural Basis of Observed Spectra. Chem. Mater. 1990, 2, 557–563. [Google Scholar] [CrossRef]
  18. Mao, J.; Gu, Q.; Gregory, D.H. Revisiting the Hydrogen Storage Behavior of the Na–O–H System. Materials 2015, 8, 2191–2203. [Google Scholar] [CrossRef]
  19. Bahranowski, K.; Klimek, A.; Gawel, A.; Serwicka, E.M. Rehydration Driven Na-Activation of Bentonite—Evolution of the Clay Structure and Composition. Materials 2021, 14, 7622. [Google Scholar] [CrossRef]
  20. Liu, S.; Dou, Z.; Zhang, S.; Zhang, H.; Guan, X.; Feng, C.; Zhang, J. Effect of Sodium Hydroxide on the Carbonation Behavior of β-Dicalcium Silicate. Constr. Build. Mater. 2017, 150, 591–594. [Google Scholar] [CrossRef]
  21. Siriwardane, R.V.; Poston, J.A.; Robinson, C.; Simonyi, T. Effect of Additives on Decomposition of Sodium Carbonate: Precombustion CO2 Capture Sorbent Regeneration. Energy Fuels 2011, 25, 1284–1293. [Google Scholar] [CrossRef]
  22. Abid, F.; Bagreev, A.; Bandosz, T.J. Effect of Surface Characteristics of Wood-Based Activated Carbons on Adsorption of Hydrogen Sulfide. J. Colloid Interface Sci. 1999, 214, 407–415. [Google Scholar] [CrossRef]
  23. Abid, M.; Garcia, R.; Martinez-Escandell, M.; Fullana, A.; Silvestre-Albero, J. Exceptional Performance of Fe@Carbon-Rich Nanoparticles Prepared via Hydrothermal Carbonization of Oil Mill Wastes for H2S Removal. Chemosphere 2024, 358, 142140. [Google Scholar] [CrossRef]
  24. Peluso, A.; Gargiulo, N.; Aprea, P.; Caputo, D.; Pepe, F. Modeling Hydrogen Sulfide Adsorption on Chromium-Based MIL-101 Metal Organic Framework. Sci. Adv. Mater. 2014, 6, 164. [Google Scholar] [CrossRef]
  25. Mao, D.; Griffin, J.M.; Dawson, R.; Fairhurst, A.; Gupta, G.; Bimbo, N. Porous Materials for Low-Temperature H2S-Removal in Fuel Cell Applications. Sep. Purif. Technol. 2021, 277, 119426. [Google Scholar] [CrossRef]
  26. Zhao, Y.; Hu, G. Removal of Sulfur Dioxide from Flue Gas Using the Sludge Sodium Humate. Sci. World J. 2013, 2013, 573051. [Google Scholar] [CrossRef] [PubMed]
  27. Yadav, R.; Singh, G.; Mishra, A.; Verma, V.; Khan, A.; Pal, N.; Sinha, A.K. A Density Functional Theory and Experimental Study of CO2 Photoreduction to Methanol Over α-Sulfur–TiO2 Composite. Electrocatalysis 2021, 12, 56–64. [Google Scholar] [CrossRef]
  28. Lacey, M.J.; Yalamanchili, A.; Maibach, J.; Tengstedt, C.; Edström, K.; Brandell, D. The Li–S Battery: An Investigation of Redox Shuttle and Self-Discharge Behavior with LiNO3-Containing Electrolytes. RSC Adv. 2016, 6, 3632–3641. [Google Scholar] [CrossRef]
  29. Yang, C.; Suo, L.; Borodin, O.; Wang, F.; Sun, W.; Gao, T.; Fan, X.; Hou, S.; Ma, Z.; Amine, K.; et al. Unique Aqueous Li-Ion/Sulfur Chemistry with High Energy Density and Reversibility. Proc. Natl. Acad. Sci. USA 2017, 114, 6197–6202. [Google Scholar] [CrossRef]
Figure 1. (a) Nitrogen adsorption/desorption isotherms for activated carbon RG before and after the incorporation of different amounts of NaOH and (b) pore size distribution after application of the QSDFT model (slit/cylinder/sphere pores; adsorption branch).
Figure 1. (a) Nitrogen adsorption/desorption isotherms for activated carbon RG before and after the incorporation of different amounts of NaOH and (b) pore size distribution after application of the QSDFT model (slit/cylinder/sphere pores; adsorption branch).
Carbon 11 00068 g001
Figure 2. The Raman spectra (a) and XRD patterns (b) of the initial adsorbents. Inset: amplification of the Raman spectra in the 2000–3750 cm−1 region.
Figure 2. The Raman spectra (a) and XRD patterns (b) of the initial adsorbents. Inset: amplification of the Raman spectra in the 2000–3750 cm−1 region.
Carbon 11 00068 g002
Figure 3. FE-SEM images of the carbon samples (RG-NaOH-X) of different magnifications.
Figure 3. FE-SEM images of the carbon samples (RG-NaOH-X) of different magnifications.
Carbon 11 00068 g003
Figure 4. Thermogravimetric profiles for the different NaOH-impregnated samples up to 1000 °C in a nitrogen atmosphere (solid line: weight loss; dashed line: derivative weight loss).
Figure 4. Thermogravimetric profiles for the different NaOH-impregnated samples up to 1000 °C in a nitrogen atmosphere (solid line: weight loss; dashed line: derivative weight loss).
Carbon 11 00068 g004
Figure 5. H2S breakthrough column profiles (a) and the total adsorption capacity at saturation (b) for the different samples evaluated.
Figure 5. H2S breakthrough column profiles (a) and the total adsorption capacity at saturation (b) for the different samples evaluated.
Carbon 11 00068 g005
Figure 6. (a) H2S breakthrough column profile for sample RG-NaOH-30 under different relative humidity conditions and (b) the total adsorption capacity at saturation.
Figure 6. (a) H2S breakthrough column profile for sample RG-NaOH-30 under different relative humidity conditions and (b) the total adsorption capacity at saturation.
Carbon 11 00068 g006
Figure 7. FE-SEM images of the carbon samples (RG-NaOH-X) of different magnifications after use in the H2S breakthrough column experiments.
Figure 7. FE-SEM images of the carbon samples (RG-NaOH-X) of different magnifications after use in the H2S breakthrough column experiments.
Carbon 11 00068 g007
Figure 8. XPS analysis of the RG-NaOH-30-af adsorbent after being exposed to H2S until saturation.
Figure 8. XPS analysis of the RG-NaOH-30-af adsorbent after being exposed to H2S until saturation.
Carbon 11 00068 g008
Table 1. pH of the different samples before and after the H2S column experiments.
Table 1. pH of the different samples before and after the H2S column experiments.
SamplepH-before H2S ExperimentpH-after H2S Experiment
RG7.007.06
RG-NaOH-1010.343.46
RG-NaOH-1510.364.06
RG-NaOH-2010.426.39
RG-NaOH-3010.556.62
RG-NaOH-4010.578.72
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

Abid, M.; Martínez-Escandell, M.; Silvestre-Albero, J. NaOH-Modified Activated Carbon Materials for Hydrogen Sulfide Removal. C 2025, 11, 68. https://doi.org/10.3390/c11030068

AMA Style

Abid M, Martínez-Escandell M, Silvestre-Albero J. NaOH-Modified Activated Carbon Materials for Hydrogen Sulfide Removal. C. 2025; 11(3):68. https://doi.org/10.3390/c11030068

Chicago/Turabian Style

Abid, Meriem, Manuel Martínez-Escandell, and Joaquín Silvestre-Albero. 2025. "NaOH-Modified Activated Carbon Materials for Hydrogen Sulfide Removal" C 11, no. 3: 68. https://doi.org/10.3390/c11030068

APA Style

Abid, M., Martínez-Escandell, M., & Silvestre-Albero, J. (2025). NaOH-Modified Activated Carbon Materials for Hydrogen Sulfide Removal. C, 11(3), 68. https://doi.org/10.3390/c11030068

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop