Valentina
Ospina-Montoya
a,
Samuel
Aguirre-Contreras
c,
Raúl
Ocampo-Pérez
c,
Erika
Padilla-Ortega
c,
Sebastián
Pérez
d,
Juan
Muñoz-Saldaña
d,
Jazmín
Porras
*b,
Nancy
Acelas
*a and
Angélica
Forgionny
*a
aGrupo de investigación Materiales con Impacto (Mat&mpac), Facultad de ingenierías, Instituto de Ciencias Básicas, Universidad de Medellín, Carrera 87 No. 30-65, Medellín 050026, Colombia. E-mail: nyacelas@udemedellin.edu.co; mforgionny@udemedellin.edu.co; Tel: +57(604) 5904500
bGrupo de Investigaciones Biomédicas UniRemington, Corporación Universitaria Remington, Calle 51 No. 51-27, Medellín 050010, Colombia. E-mail: jazmin.porras@uniremington.edu.co; Tel: +57(604)3221000
cCentro de Investigación y Estudios de Posgrado, Facultad de Ciencias Químicas, Universidad Autónoma de San Luis Potosí, San Luis Potosí, 78260, Mexico
dCentro de Investigación y de Estudios Avanzados del Instituto Politecnico Nacional, Laboratorio Nacional de Proyección Térmica (CENAPROT), Libramiento Norponiente 2000 Fracc, Real de Juriquilla, 76230, Querétaro, Mexico
First published on 17th September 2025
This study elucidates the competitive adsorption dynamics of ciprofloxacin (CIP) and acetaminophen (ACE) onto coffee husk activated with potassium (CH–KOH, BET surface area = 1145 m2 g−1, pHPZC = 7.36), providing mechanistic insights into the removal of pharmaceuticals in complex aqueous matrices. The Modified Langmuir multicomponent isotherm effectively captured the competitive equilibrium behavior (deviation = 25.8%), showing a higher affinity for ACE (ηACE = 0.7) than for CIP (ηCIP = 4.9), the qT was 1.25 mmol g−1 across the entire evaluated concentration range, which is similar to the observed in mono-component systems, Qmax 1.26 mmol g−1 for ACE and 0.58 mmol g−1 for CIP, with removal efficiencies of 91–99% and 75–99%, respectively. In real matrices such as synthetic hospital wastewater and urine, high efficiencies (84–97%) were maintained. Fixed-bed column studies confirmed the strong performance under continuous-flow conditions, with saturation capacities (qs) up to 1.46 mmol g−1 for ACE and 0.61 mmol g−1 for CIP, mass transfer zones ranging from 0.42 to 1.53 cm, and breakthrough times between 91 and 1463 min depending on flow rate (1–3 mL min−1) and bed height (1–3 cm). The Thomas model accurately predicted breakthrough curves, revealing faster kinetics for ACE. Physisorption predominates, involving synergistic π–π stacking interactions, hydrogen bonding networks, and hydrophobic association, with ACE showing greater selectivity in both mono and multicomponent systems. CH–KOH exhibited high stability and reusability, stabilizing at approximately 70% of its initial capacity by the third cycle, with no further decrease observed in the fourth cycle. Comprehensive physicochemical characterization revealed that physisorption predominates, involving synergistic π–π stacking interactions, hydrogen bonding networks, and hydrophobic associations. These results confirm the potential of CH–KOH as a sustainable adsorbent for pharmaceutical contaminant removal in real-world scenarios, integrating circular economy principles into advanced water treatment.
Water impactThis study valorizes coffee husk waste into an efficient adsorbent for removing pharmaceuticals from complex wastewater. By demonstrating high performance in continuous-flow systems and realistic matrices (synthetic hospital wastewater and urine), the work highlights a scalable circular economy solution for sustainable water treatment. |
Acetaminophen (ACE) and ciprofloxacin (CIP) exemplify the challenges posed by EPs in wastewater systems. ACE, a commonly consumed analgesic, is often detected in effluents, with 58–68% of ingested doses excreted unchanged in urine.11 While ACE is moderately toxic to aquatic organisms, its degradation byproducts (e.g., 4-aminophenol) are mutagenic and carcinogenic.12 CIP, a fluoroquinolone antibiotic that poses additional risks, including toxicity to aquatic life and the promotion of antibiotic-resistant bacteria.13 Despite these threats, regulatory frameworks for controlling their discharge remain inadequate in many regions.
Conventional wastewater treatment plants (WWTPs), which primarily rely on physical and biological processes, are not specifically designed to eliminate micropollutants such as pharmaceuticals. These compounds often exhibit low biodegradability and strong stability under standard treatment conditions, leading to incomplete removal. In some cases, effluent concentrations of certain drugs have been reported to be higher than influent levels. This may be due to the desorption of compounds from sludge, biotransformation of conjugates back to the parent drug, or analytical matrix suppression in influents. Such behavior has been observed in various studies, Botero-Coy et al. (2018).14–18 Advanced technologies, such as electrocoagulation,19 sonochemical treatment,20 biological treatment,20 advanced oxidation,21 membrane filtration processes,22,23 and photocatalytic degradation24 have shown promise, but their high energy demands, operational costs, and secondary waste generation limit scalability. In contrast, adsorption using activated carbon is a cost-effective, efficient alternative that leverages its high surface area, porosity, and chemical stability.25–29
Current research on removing pharmaceuticals from wastewater predominantly focuses on mono-component batch systems, failing to address the competitive adsorption dynamics characteristic of real wastewater matrices. Comprehensive multicomponent analyses and continuous-flow experiments are essential to bridge this critical knowledge gap for optimizing operational parameters in practical applications. Selecting appropriate adsorbent materials is another crucial consideration in this field. Although activated carbon has excellent adsorption capacity, its high production costs and disposal challenges significantly limit large-scale implementation.30
In response to these limitations, agro-industrial waste materials have emerged as sustainable and low-cost alternatives for producing biochar and activated carbon with optimal adsorption properties. Numerous studies have demonstrated the effectiveness of various agricultural byproducts for this purpose, including sugar cane bagasse,31 coffee waste,32 spent tea leaves,33 grass,34 beer grains,35 avocado seeds,36 bamboo chips,37 cellulose factory sludge38 in removing pharmaceutical compounds.
Chemical activation processes using agents such as phosphoric acid, potassium hydroxide (KOH) or zinc chloride significantly enhance the porosity and pore volume structure of these materials. Among these activating agents, KOH has proven particularly effective, consistently producing adsorbents with well-defined pore structures, a high specific surface area and favorable yields.31,33,35,37
Recent investigations have confirmed the exceptional performance of KOH-activated agricultural waste-derived carbons in removing EPs, positioning them as promising sustainable solutions for advanced wastewater treatment.39–42
Egbedina et al. (2021)65 prepared KOH-activated coconut husk–kaolinite biochar via a microwave technique, achieving a Qmax 229 mg g−1 for CIP in batch mode. Liu et al. (2023)66 reported activated carbon from waste peony seed shells with an ultra-high specific surface area (2980.96 m2 g−1) and CIP removal (Qmax = 782.3 mg g−1). In contrast, Li et al. (2018)67 obtained potato stem–leaf biochar (38.75 m2 g−1) activated with KOH, which reached 23.36 mg g−1 for CIP removal. In these cases, adsorption mechanisms included hydrophobic interactions, hydrogen bonding, electrostatic attraction, and π–π interactions.
The accumulation of environmental impact of agro-industrial waste presents a particularly pressing challenge, in coffee-producing nations like Colombia. As the world's third-largest coffee producer after Brazil and Vietnam, Colombia harvested 820600 tons of coffee from 839
700 hectares in recent years, generating substantial waste volumes.43,44 The coffee processing chain remains remarkable inefficient, generating waste that ranges from 30% to 50% of the total weight of coffee processed.45 Specifically, processing one ton of fresh coffee cherries yields approximately 0.5 tons of pulp and 0.18 tons of husk.46 Currently, these residues lack economically viable applications and often accumulate in landfills, creating serious environmental concerns.46,47 The high lignin content (21%) and slow degradation rates of coffee processing byproducts exacerbate these issues, leading to methane emissions and other negative environmental impacts when they accumulate in stockyards and dumpsites.48,49 Converting coffee husks into value-added adsorbent materials offers a dual solution by simultaneously addressing waste management challenges in the coffee industry and providing sustainable water treatment options that align with circular economy principles.
This study aims to advance understanding of the adsorption process of ciprofloxacin and acetaminophen adsorption onto KOH-activated coffee husk-derived carbon through a comprehensive experimental approach.
Therefore, this work focuses on exploring the adsorption behavior of these compounds in greater detail using various experimental techniques. The methodology progresses systematically from fundamental batch experiments to practical applications, beginning with mono-component isotherm studies to establish baseline adsorption characteristics. Interactions between the adsorbent and contaminants were analyzed using advanced spectroscopic techniques, including Fourier transform infrared spectroscopy (FTIR) and X-ray photoelectron spectroscopy (XPS). The investigation then advances to more realistic scenarios through multicomponent analysis and continuous-flow fixed-bed column experiments specifically targeting CIP and ACE removal. Additionally, the adsorbent's recyclability was evaluated alongside its performance in complex matrices such as synthetic wastewater and urine—an approach rarely reported for agricultural waste-derived carbons to assess practical feasibility. By integrating fundamental adsorption studies with fixed-bed continuous-flow evaluations under realistic water matrices, this work bridges the gap between lab-scale testing and real-world application, providing critical insights for developing improved contaminant removal strategies in complex wastewater systems.
The surface functional groups were characterized using Fourier transform infrared spectroscopy with attenuated total reflectance (ATR-FTIR) accessory, scanning the spectral range of 4000 to 450 cm−1 with a Spectrum Two spectrometer (PerkinElmer, Waltham, Massachusetts, USA). The phase structure of CH–KOH was analyzed using X-ray diffraction (XRD) with a Rigaku SmartLab diffractometer operating at 44 kV and 40 mA and equipped with a Cu Kα1 radiation source (λ = 1.5406 Å) both before and after adsorption. Diffraction patterns were recorded over a 2θ range of 10° to 70°, with a step size of 0.02° and a scan rate of 5° min−1. The microstructure of CH–KOH, before and after adsorption was characterized using a JSM-7610F scanning electron microscope (SEM, JEOL) operating at an acceleration voltage of 15 kV, utilizing a secondary electron detector for imaging.
The surface chemical composition and oxidation states of CH–KOH, before and after adsorption were analyzed using X-ray photoelectron spectroscopy (XPS). Spectra were recorded using a Scanning XPS micro – probe PHI 5000 VersaProbe II, with an Al K X-ray source (hν = 1486.6 eV) monochromatic with a 100 μm beam diameter, and an MCD analyzer. The binding energy (BE) scale was calibrated by setting the C1s peak at 284.8 eV. Spectral deconvolution was performed by curve fitting following Shirley-type background subtraction and asymmetric functions were considered with a line shape SGL(p)T(k) utilizing the CasaXPS software.55
The contaminant solutions were introduced using a peristaltic pump (SHENCHEN LabV6-111) at controlled flow rates. Contaminant concentrations were measured at various time intervals from the beginning of the experiment until column saturation (C/C0 = 1 or constant), and the breakthrough curves were plotted. A series of characteristics of the adsorbent in the continuous flow column were determined using breakthrough curves. These characteristics included saturation time (ts), breakthrough time (tb), mass transfer zone (MTZ), the adsorption capacity of the column at the breakthrough time (qb), and adsorption capacity of the column at the saturation time (qs) calculated using eqn (S12) and (S13). The experimental rupture curves were fitted to two common models for interpretation, represented by Thomas60 (eqn (S14)) and Clark61 (eqn (S15)).
The reusability of CH–KOH was assessed through four consecutive adsorption–desorption cycles. The adsorbent was recovered by vacuum filtration after adsorption and then regenerated via ultrasonic treatment (Digital Pro, PS-30AL) at 40 KHz for 30 min in a solution containing 33.3 mL methanol and 6.6 mL of 3% NaOH solution. The regenerated material was then vacuum washed with 700 mL deionized water and dried at 100 °C for 24 h.
On the other hand, ACE (pKa 9.5) demonstrates limited electrostatic interactions at pH values below 10, because it is in its neutral form. The consistent removal efficiency across various pH conditions can be attributed to multiple adsorption mechanisms not pH-dependent, including pore-filling processes and molecular interactions between the aromatic rings and the carbon structure of this pharmaceutical compound. Therefore, all subsequent experiments were conducted at pH = 6 ensuring optimal adsorption performance and balancing electrostatic and non-electrostatic interactions.
The analysis of the adsorption isotherms for the removal of ACE and CIP using the CH–KOH material reveals important information about the adsorption process. The experimental data were analyzed using the Langmuir and Freundlich models (see Fig. 1c and d). As shown in Table 1, the determination coefficients (R2Adj) were 0.991 for ACE and 0.979 for CIP, indicating an adequate agreement with the Freundlich model. This suggests that the adsorption process occurs mainly on a heterogeneous surface with multiple adsorption sites of different energy levels. The suitability of the Freundlich model highlights the complex nature of the adsorption mechanism, which likely involves interactions such as multilayer adsorption or surface heterogeneity, consistent with the physicochemical properties of the adsorbent material. These results are related to previous studies where similar behaviors have been reported for adsorption processes on heterogeneous surfaces.62
Langmuir | Freundlich | |||||
---|---|---|---|---|---|---|
Q max (mmol g−1) | K L (L g−1) | R 2Adj | K F (mmolg−1 (L mg−1)−1/n) | n F | R 2Adj | |
CH–KOH–ACE | 1.26 ± 0.06 | 194 ± 50 | 0.951 | 2.3 ± 0.1 | 0.24 ± 0.01 | 0.991 |
CH–KOH–CIP | 0.58 ± 0.01 | 863 ± 186 | 0.992 | 0.66 ± 0.04 | 0.06 ± 0.02 | 0.979 |
The maximum adsorption capacity, according to the Langmuir model, Qmax was significantly higher for ACE (1.26 mmol g−1) compared to CIP (0.58 mmol g−1). The selectivity ratio, S = (Qmax,ACE/Qmax,CIP), derived from the individual adsorption isotherms, indicates that CH–KOH has an ACE adsorption capacity 2.2 times greater than CIP. This suggests that CH–KOH exhibits a higher affinity for ACE molecules in a mono-component system. This behavior can be explained by the different molecular characteristics of the two contaminants and their interaction with the material surface. The CIP adsorption process is highly favored by the occurrence of different chemical interactions. In contrast, the ACE adsorption process is governed by physical adsorption related to a high surface area of CH–KOH material and, to a lesser extent, to the occurrence of chemical interactions between ACE molecules and the carbonaceous material, because of the neutral form of ACE to the evaluated conditions. The values of the nF parameter of the Freundlich model, which were 0.24 and 0.06 for ACE and CIP, respectively, are less than 1, indicating that adsorption is favorable for both compounds. The lower nF value for CIP suggests greater heterogeneity in the adsorption process of this compound, which could be related to different interaction mechanisms between the CIP molecule and the adsorbent surface.63–65
In comparison with literature-reported adsorbents derived from agricultural wastes (see Table S7), CH–KOH shows competitive adsorption capacities.
For example, KOH-activated coconut husk–kaolinite biochar prepared via a microwave technique has been reported to reach a qm of 0.69 mmol g−1 (229 mg g−1) for CIP in batch mode.65 Activated carbon derived from waste biomass such as peony seed shells, with an ultra-high specific surface area (2980.96 m2 g−1), achieved CIP removal of 2.36 mmol g−1 (Qmax = 782.3 mg g−1) in simplified aqueous matrices.66 Biochar from potato stems and leaves, prepared by pyrolysis at 500 °C and activated with KOH (38.75 m2 g−1), exhibited a qm of 0.07 mmol g−1 (23.36 mg g−1). In these cases, hydrophobic interactions, hydrogen bonding, electrostatic attraction, and π–π interactions were identified as the main adsorption mechanisms for CIP uptake.67 In the present study, CH-KOH achieved qm values of 0.58 mmol g−1 for CIP and 1.26 mmol g−1 for ACE, maintaining high performance even under multicomponent and continuous-flow conditions.
The LME and MLMηi models were applied to analyze the multicomponent adsorption data and demonstrated satisfactory agreement with the experimental findings. The LME and MLMηi models exhibited deviation percentages (DESV) of 25.3% and 25.8%, respectively.
Although the LME model showed slightly lower deviation, it is not considered optimal because all of its parameters are determined by fitting competitive adsorption equilibrium data without incorporating individual Langmuir isotherm parameters. In contrast, the MLMηi model considers the parameters of each isotherm and includes an interaction factor specific to each species. Despite the moderate fitting deviation, MLMηi provided a more realistic representation of the system compared to traditional isotherm models, supporting its relevance as a first-step predictive tool for complex adsorption processes on heterogeneous surfaces.
The interaction factor values were ηACE = 0.7 and ηCIP = 4.9, suggesting that the material's surface has a significantly higher affinity for ACE than CIP in multicomponent systems. Specifically, ηCIP is seven times larger than ηACE, indicating a stronger competitive effect on CIP adsorption than on ACE adsorption. Moreover, the models of the multicomponent Freundlich isotherm were used to analyze binary adsorption equilibrium data. However, they exhibited deviations greater than 50% and were therefore not considered.
Fig. 2a and b show the MLMηi model predictions, along with the binary adsorption equilibrium data for CIP and ACE on CH–KOH. Fig. 2a shows the adsorption response surface for ACE adsorption and reveals that CIP presence affected ACE adsorption at low CIP concentrations within the evaluated range (0.04–0.16 mmol L−1). The inhibitory effect was most significant at ACE concentration of 0.007 mmol L−1, where adsorption capacity decreased by 44% compared to the individual system. At higher ACE concentrations (0.11 mmol L−1), however the reduction in adsorption capacity was less pronounced, decreasing by 25.4% relative to the ACE adsorption capacity in the individual system. For CIP adsorption (Fig. 2b), a noticeable effect of ACE presence was observed at lower ACE concentrations (0–0.06 mmol L−1). At these concentrations, the adsorption capacity of CIP decreased by 56% at 0.05 mmol L−1 ACE and 33% at 0.06 mmol L−1 ACE, respectively compared to the individual system. Additionally, a reduction of nearly 30% was observed at ACE concentrations of 0.07 and 0.15 mmol L−1. These results demonstrate an antagonistic effect and competition between the two pharmaceutical compounds for adsorption sites on CH–KOH, particularly at lower concentrations in the multicomponent system. Additionally, the competitive effects depend on the equilibrium concentration. Fig. 2c shows the total adsorbed mass in the multicomponent system (qT = qACE + qCIP) confirming that both pharmaceuticals compete for the same adsorption sites. This is supported by the fact that qT remains around 1.25 mmol g−1 across the entire evaluated concentration range, a value like the Qmax found for ACE in the mono-component system.
The competitive adsorption behavior observed in this study aligns with the findings reported by Pauletto, et al.68 regarding acetaminophen–nimesulide binary systems. The presence of the competing pharmaceutical compound significantly impacted ACE adsorption. The authors reported a 28% reduction in ACE adsorption capacity when co-adsorbed with nimesulide. They are attributed to a displacement phenomenon, whereby ACE molecules were released from active sites due to nimesulide's higher affinity for the adsorbent. However, our system exhibited a more concentration-dependent competitive effect, with the strongest inhibition observed at low ACE concentrations. These findings suggest that ACE's adsorption is susceptible to competitive effects, regardless of the competing pharmaceutical compound.
Fig. 3c shows the adsorbent's performance over multiple adsorption–desorption cycles, a critical factor in assessing the material's practical viability. Reusability is essential for both sustainability and cost-effectiveness. After the first cycle, the material maintained high effectiveness, with only a 20% reduction in removal efficiency. The adsorption capacities decreased from 1.26 to 0.85 mmol g−1 for ACE and from 0.58 to 0.40 mmol g−1 for CIP, corresponding to approximately 70% removal in both cases. From the third to the fourth cycle, no further decrease in removal capacity was observed for CH–KOH, indicating stable performance after initial use.
This moderate reduction could be attributed to partial active site saturation. The minimal decline suggests that most active sites remain available and functional for subsequent adsorption cycles. The ability of the adsorbent to maintain stable performance across multiple usage cycles demonstrates CH–KOH remarkable chemical and structural stability. This behavior can be ascribed to the unique characteristics of the CH precursor and the chemical activation process using KOH, which generated a material with a robust porous structure and high density of accessible active sites. Furthermore, resistance of the material to structural degradation or porosity collapse indicates that the activation process produced a substrate with exceptional mechanical and chemical resilience. CH–KOH's consistent removal capacity throughout multiple cycles suggests its potential applicability in continuous treatment systems or batch treatment applications, where long-term stability is essential. This stability also minimizes the need for frequent regeneration, consequently reducing operational costs and secondary waste generation. CH–KOH is an attractive candidate for large-scale implementation in advanced water treatment technologies.
Column characteristics | ||||||||
---|---|---|---|---|---|---|---|---|
Flow (mL min−1) | t b (min) | t s (min) | MTZ (cm) | q e (mmol g−1) | q s (mmol g−1) | Adsorbent mass (g) | U r (g L−1) | |
ACE | ||||||||
a) | 1 | 904.71 | 2095.17 | 0.57 | 0.71 | 1.17 | 0.12 | 0.13 |
2 | 499.75 | 1140.06 | 0.56 | 0.78 | 1.20 | 0.12 | 0.12 | |
3 | 365.04 | 625.60 | 0.42 | 0.90 | 1.23 | 0.12 | 0.11 |
CIP | ||||||||
---|---|---|---|---|---|---|---|---|
b) | 1 | 287.20 | 820.25 | 0.65 | 0.23 | 0.43 | 0.12 | 0.42 |
2 | 91.17 | 532.92 | 0.83 | 0.15 | 0.44 | 0.12 | 0.66 | |
3 | 110.65 | 544.70 | 0.80 | 0.28 | 0.61 | 0.12 | 0.36 |
Bed (cm) | t b (min) | t s (min) | MTZ (cm) | q e (mmol g−1) | q s (mmol g−1) | Adsorbent mass (g) | U r (g L−1) | |
---|---|---|---|---|---|---|---|---|
ACE | ||||||||
c) | 1 | 365.04 | 625.60 | 0.42 | 0.90 | 1.23 | 0.12 | 0.33 |
2 | 845.50 | 1227.40 | 0.62 | 1.04 | 1.26 | 0.24 | 0.14 | |
3 | 1462.64 | 2222.17 | 1.03 | 1.16 | 1.46 | 0.36 | 0.08 |
CIP | ||||||||
---|---|---|---|---|---|---|---|---|
d) | 1 | 91.17 | 532.92 | 0.83 | 0.15 | 0.44 | 0.12 | 1.32 |
2 | 370.06 | 708.36 | 0.96 | 0.27 | 0.42 | 0.24 | 0.32 | |
3 | 664.79 | 1354.50 | 1.53 | 0.37 | 0.58 | 0.36 | 0.18 |
Concentrations (mmol L−1) | t b (min) | t s (min) | MTZ (cm) | q e (mmol g−1) | q s (mmol g−1) | Adsorbent mass (g) | U r (g L−1) | |
---|---|---|---|---|---|---|---|---|
ACE | ||||||||
e) | 0.09 | 845.50 | 1227.40 | 0.62 | 1.04 | 1.26 | 0.24 | 0.09 |
0.12 | 605.03 | 1024.06 | 0.82 | 0.89 | 1.23 | 0.24 | 0.13 | |
0.15 | 429.17 | 968.37 | 1.11 | 0.80 | 1.37 | 0.24 | 0.19 |
CIP | ||||||||
---|---|---|---|---|---|---|---|---|
f) | 0.09 | 91.17 | 532.92 | 0.83 | 0.15 | 0.44 | 0.12 | 0.65 |
0.12 | 93.86 | 338.25 | 0.72 | 0.19 | 0.41 | 0.12 | 0.63 | |
0.15 | 49.64 | 253.78 | 0.80 | 0.13 | 0.36 | 0.12 | 1.19 |
Table 2 includes the usage ratio (Ur), which represents the volume of solution treated per gram of adsorbent up to the breakthrough point. For ACE, Ur values ranged from 0.08 g L−1 to 0.33 g L−1, whereas for CIP, Ur values ranged from 0.18 g L−1 to 1.32 g L−1. These results clearly indicate that, under continuous operation, the removal of CIP is more efficient than that of ACE.
Spectroscopic characterization techniques were employed to confirm the interactions between CH–KOH and the contaminants. Fig. 6 shows: a) the FTIR spectrum in the range of 3950–3000 cm−1, b) the FTIR spectra deconvolution in the range of 1700–1300 cm−1, c) the FTIR spectra deconvolution in the range of 1300–450 cm−1, d) the second derivative of the FTIR spectra, which was used to identify the main functional group bands during the deconvolution process, and e) XRD patterns, which confirm the presence of elements identified by SEM–EDS. Additionally, difference spectra before and after adsorption, presented in Fig. S5, were employed to identify new band positions and the reduction of specific peaks. The FTIR spectra reveal characteristic O–H group signals at 3384 and 3206 cm−1, indicating the presence of these functional groups before and after adsorption. In the 1700–1300 cm−1 region, CH–KOH exhibits signals at 1619 cm−1, 1580 cm−1, 1548 cm−1, and 1507 cm−1, which can be assigned to aromatic (Ar) CC, and reflect the presence of diverse aromatic structures. The band at 1471 cm−1 can be assigned to υC–O and –CH2, which aligns with the structural complexity typically found in activated carbon materials.71 Signals in the 1300–450 cm−1 range confirm oxygen-containing functional groups, with peaks at 1214 cm−1 (Ar–C–O), 1150 cm−1 (C–O, ester), 1074 cm−1 (C–OH), 1020 cm−1, and 957 cm−1 (C–O).54 The CH–KOH–ACE sample shows new bands at 1576 cm−1, which correspond to C
C–N–H bending (amide); at 1540 cm−1 for C–N–C
O stretching (amide); at 1500 cm−1 for aromatic C
C, 1481 cm−1 assigned to –CH3 antisymmetric bending; at 1465 cm−1 for O–H in-plane bending; at 802 cm−1 for C–H aromatic out-of-plane bending, and 769 cm−1 for C
C. Additionally, the C–OH band at 1074 cm−1 in the CH–KOH sample weakens, and a new band appears at 1028 cm−1. These changes indicate strong interactions between the phenolic groups of CH–KOH and ACE, consistent with the structural behavior of ACE at higher pH levels (see Fig. 1b). In these conditions, the phenolic–OH group contributes to the FTIR band shift from 1074 to 1028 cm−1.72 After CIP adsorption, however, the CH–KOH–CIP spectrum in the 1700–1300 cm−1 range shows new signals at 1607 cm−1, which correspond to C
O bond vibrations or quinolone N–H bending vibration. In the 1300–1500 cm−1 range, new peaks appear at 1562 cm−1, 1502 cm−1, and 1483 cm−1 for aromatic C
C, along with bands at 1450 cm−1 for υC–O and 1380 cm−1 for C–H. Additionally, signals are observed at 1230 cm−1, which are assigned to C–O in O
C–O, and at 888 cm−1 and 801 cm−1 C
C and C–H respectively for the aromatic structure.73 These bands confirm the presence of CIP.
The XRD patterns (Fig. 6e, and Table S8), CH–KOH show two broad Bragg peaks at 22.0° and 43.7°, which correspond to the (002) and (100) planes, respectively. These peaks are characteristic of the hexagonal graphite structure. These broad peaks are typical of amorphous carbon materials.74 Additionally, two low-intensity peaks were identified: one at 2θ = 21.6°, which was assigned to the (111) plane of SiO2 (ICSD 162659) with a 1.1 wt% according to the Rietveld refinement, and another at 2θ = 29.6°, which was associated with the (11−2) plane of CaCO3 (ICSD collection code: 150) with a 0.7 wt%. After the adsorption of ACE and CIP, the peak corresponding to CaCO3 was no longer observed. Nevertheless, the material continues to exhibit the characteristic peaks for the (002) and (100) planes of the carbonaceous structure of activated carbon, No CaCO3 was present, and only 0.7 wt% of SiO2 was detected in CH–KOH–CIP.
Fig. 7 and Table S9 present the XPS analysis of (a) C1s, (b) O1s, (c) N1s, and (d) F1s before and after the adsorption of ACE and CIP, revealing significant changes in the signals corresponding to the main elements (C, O, N, and F). In the CH–KOH sample, the following C1s contributions were identified: CC (60.3 at%), C–C (15.0 at%), C–OH (14.2 at%), C
O (6.4 at%), and O
C–O (4.1 at%). After adsorption, these values changed to C
C (48.6 at%), C–C (21.2 at%), C–OH (19.5 at%), C
O (7.2 at%), and O
C–O (3.4 at%), indicating an increase in C–OH and C
O groups.74
For O1s, prior to adsorption, peaks were observed at 531.8 eV for aromatic OC (27.7 at%), 533.1 eV for O
C–OH (30.7 at%), 534.2 eV for O–H in O
C–OH (32.4 at%), and 535.5 eV for adsorbed water (9.2 at%). After adsorption, a shift in the O
C–OH signal was observed around 535 eV (39.1 at%), along with the appearance of a new peak at 533.30 eV assigned to C–O (8.9 at%), both representing oxygen in O–C and O
C within the O
C–OH (11.6 at%) group. Additionally, a new signal at 531.8 eV was detected and attributed to oxygen in the O
C–N (14.8 at%) group of ACE.
Initially, the N1s spectrum of CH–KOH displayed peaks at 399.7 eV (37.4 at%) for pyrrolic nitrogen (Pyr), 401.7 eV (19.0 at%) for graphitic nitrogen (Gra), and 404.2 eV (25.0 at%) for NOx species. After adsorption, new bands appeared at 400.1 eV (35.2 at.%), attributed to the C–NR2 group, and at 401.3 eV (25.4 at%), corresponding to the OC–NHR group for ACE and CIP, respectively.
Finally, the F1s spectra exhibited peaks at 687.5 eV (51.3 at%) and 690.4 eV (48.7 at%). Although small, the signal at 690.4 eV suggests an interaction between fluorine and the activated carbon surface.75
Fig. 8 summarizes the main mechanisms identified in the adsorption of ACE and CIP onto CH–KOH. Based on information obtained from the adsorption experiments and material characterization, the adsorption mechanism is primarily governed by physisorption, involving non-covalent interactions such π+–π interactions, hydrophobic interactions, hydrogen bonding, and van der Waals forces.
![]() | ||
Fig. 8 Various mechanisms of CH–KOH and interaction with acetaminophen (ACE) and ciprofloxacin (CIP). |
For ACE, the molecules remain nearly neutral across the evaluated pH values. Non-electrostatic interactions may include hydrogen bonding, π–π interactions, n–π electron donor–acceptor interactions, and hydrophobic–hydrophobic interactions. CH–KOH has various polar surface functionalities, such as –OH and –COOH groups, that can interact with the –OH and –NH moieties in ACE molecules through hydrogen bonding. This was evidenced by the differences observed in the FTIR spectra. ACE adsorption on the CH–KOH surface can also occur via π–π interactions between the π electrons of the fused conjugated aromatic moieties in CH–KOH (donor) and the π electrons in the benzene ring of ACE (acceptor). ACE adsorption can also occur through n–π electron donor–acceptor interactions between the lone pair of electrons on the oxygen in the surface carbonyl groups of CH–KOH (electron donors) and the benzene ring of ACE (electron acceptor). The significant contribution of the π–π mechanism to ACE sorption is further supported by the adsorption capacity remaining nearly unchanged under varying pH conditions, as shown in Fig. 1, given that π–π interactions are pH-independent. Additionally, pore diffusion is expected to play a substantial role in the uptake of ACE.
The multicomponent adsorption data were described by the MLMηi model, indicating significantly higher affinity for ACE (ηACE = 0.7) than for CIP (ηCIP = 4.9), with a qT of 1.25 mmol g−1 like the Qmax found for ACE in the mono-component system. Competitive effects were concentration-dependent, with ACE reducing CIP adsorption by up to 44% and CIP reducing ACE adsorption by up to 25%.
The potential for application in real treatment systems is reinforced by the systematic evaluation in multicomponent batch systems and in continuous fixed-bed columns. In columns, CH–KOH reached saturation capacities (qs) up to 1.46 mmol g−1 for ACE and 0.61 mmol g−1 for CIP, with mass transfer zones of 0.42–1.53 cm and breakthrough times from 91 to 1463 min depending on flow rate (1–3 mL min−1) and bed height (1–3 cm). The Thomas model successfully described the breakthrough behavior, confirming faster kinetics for ACE.
Physisorption predominates, involving synergistic π–π stacking interactions, hydrogen bonding networks, and hydrophobic association, supported by detailed characterization before and after adsorption (FTIR, XRD, XPS, SEM–EDS), which confirmed the interactions between the adsorbent and contaminants and verified the stability of the material's structure.
Reusability studies showed that CH–KOH maintained ∼70% removal capacity after three adsorption–desorption cycles, and no further gradual loss in adsorption capacity beyond the third cycle, demonstrating its long-term stability. The combination of biomass valorization, enhanced adsorption properties from KOH activation, strong performance in complex matrices, detailed mechanistic understanding, and operational stability highlights CH–KOH as a promising adsorbent for large-scale pharmaceutical contaminant removal, fully aligned with sustainable and circular economy strategies.
The authors confirm that the data supporting the findings of this study are available within the article.
This journal is © The Royal Society of Chemistry 2025 |