Open Access Article
Zenaida Guerra-Que
*a,
Katia S. López-Margalli
a,
Juan Manuel Urrieta-Saltijerala,
Adib Abiu Silahua-Pavóna,
Héctor Martínez-Garcíaa,
Pedro García-Alamillab,
Gerardo E. Córdova-Pérezc,
Juan Carlos Arévalo-Pérez
d and
José Gilberto Torres-Torresd
aTecnológico Nacional de México-Instituto Tecnológico de Villahermosa (TECNM – I. T. Villahermosa), Laboratorio de Investigación 1 Área de Nanotecnología, Km. 3.5 Carretera Villahermosa–Frontera, Cd. Industrial, C.P. 86010 Villahermosa, Tabasco, Mexico. E-mail: zenaida.gq@villahermosa.tecnm.mx; zenaida.guerra4@gmail.com; Tel: +52 9933046111
bDivisión Académica de Ciencias Agropecuarias (DACA), Universidad Juárez Autónoma de Tabasco (UJAT), Carret. Villahermosa-Teapa Km 25, Ra. La Huasteca, Tabasco, 86280, Mexico
cBenemérita Universidad Autónoma de Puebla, Facultad de Ciencias químicas, Centro Avanzado de Pruebas Analíticas no Destructivas, 72570 Puebla, Mexico
dUniversidad Juárez Autónoma de Tabasco, Centro de Investigación de Ciencia y Tecnología Aplicada (CICTAT), DACB, Laboratorio de Nanomateriales Catalíticos Aplicados al Desarrollo de Fuentes de Energía y Remediación Ambiental, Km.1 carretera Cunduacán-Jalpa de Méndez, Cunduacán, Tabasco, C.P. 86690, Mexico
First published on 1st December 2025
This study reports the green synthesis of activated carbon (ANC) from Cocoa Pod Husk (CPH), a lignocellulosic agrowaste, using sulphuric acid and sodium hydroxide treatments for Methylene Blue (MB) removal. Among the synthesised samples, ANCSH5 (NaOH, 5 M) exhibited superior performance due to extensive delignification, enhanced graphitization, and an abundance of oxygenated functionalities, as confirmed by Fourier Transform Infrared (FTIR), Scanning Electron Microscopy – Energy Dispersive X-ray (SEM–EDX), the Brunauer–Emmett–Teller Method (BET), X-ray Diffraction (XRD), Point of Zero Charge (pHPZC), and Raman analyses. ANCSH5 achieved 92% MB removal (1.16 mg g−1), surpassing CPH (24%, 0.30 mg g−1) and ANCSA10 (44%, 0.50 mg g−1). Adsorption followed pseudo-second-order kinetics, while equilibrium data fitted best to the Freundlich, Sips, Toth, and Redlich–Peterson models (R2 > 0.999). Process optimisation via a central composite design (CCD) and Response Surface Methodology (RSM) determined optimal conditions (49.6 °C, pH 6.0), achieving a 99.05% removal rate. Importantly, isotherm modelling predicted a competitive qmax of 100 mg g−1 for ANCSH5 under practical conditions. Unlike conventional biomass carbons, the hemicellulose-rich CPH precursor promoted extensive delignification, partially graphitised domains, and tailored surface chemistry, providing a distinctive adsorption mechanism. These features underscore CPH valorisation as a sustainable route to high-performance, low-cost adsorbents for dye remediation in cocoa-producing regions.
Methylene blue (MB) is a widely used cationic dye in textile and paper processing, whose release into aquatic environments can cause carcinogenic, mutagenic, and ecotoxic effects.9–11 Due to its complex aromatic structure (Fig. 1), MB is resistant to conventional treatment methods (biological or physicochemical), which are often inefficient or economically unsustainable.12,13 In this context, adsorption onto Activated Carbon (AC) has emerged as a cost-effective and efficient alternative compared to other technologies such as ion exchange, membrane filtration, or photocatalysis.13–15
In developing countries such as Mexico, wastewater from the textile industry is often discharged directly into rivers and other natural water bodies, either untreated or insufficiently treated.16,17 This practice significantly contributes to the deterioration of aquatic ecosystems and highlights the urgent need for cost-effective and sustainable wastewater treatment technologies.
Cocoa Pod Husk (CPH), an abundant agro-industrial waste in cocoa-producing regions such as Mexico, represents a largely underutilised precursor for carbon production. For every ton of dry cocoa beans, approximately 10 tons of CPH are discarded, generating a significant environmental burden.18–24 According to the Mexican Agriculture Public Organization (MAPO, 2022),25 about 28
000 tonnes of cocoa beans were produced in 2021, corresponding to nearly 280
000 tonnes of CPH waste. The lignocellulosic composition of CPH, rich in hemicellulose, lignin, and cellulose, makes it an attractive candidate for chemical activation and valorisation into carbon-based nanomaterials.18–21 Unlike other biomass sources,26–32 the high hemicellulose fraction of CPH facilitates delignification, enabling the formation of nanocarbon domains with distinct surface functionalities.
In particular, CPH has been reported as a promising feedstock in recent studies, confirming its potential for removing organic pollutants from water.21,23,33,34 The development of carbon-based materials has also been highlighted in other contexts, such as polymeric composites and structural applications, demonstrating the versatility and broad applicability of carbon materials.35 Nevertheless, limitations remain regarding pore efficiency under variable operating conditions. Moreover, few studies have applied robust statistical design approaches such as Response Surface Methodology (RSM) to optimise critical process parameters simultaneously.
In this work, we applied two independent chemical activation strategies, alkaline (NaOH) and acidic (H2SO4), to CPH, a non-conventional lignocellulosic precursor, to synthesise Activated Carbon (ANC). Unlike conventional biomass precursors such as rice husk or coconut shell, CPH possesses a uniquely high hemicellulose-to-lignin ratio, which enables more extensive delignification and the development of partially graphitised domains enriched with oxygen-containing functionalities.18,20,36,37 These distinctive structural and chemical transformations differentiate CPH-derived ANC from previously reported biomass-based carbons, providing higher graphitization and enhanced adsorption performance.38–41 The resulting materials were comprehensively characterised to elucidate surface chemistry and textural evolution, while adsorption toward MB was evaluated through kinetic and equilibrium modelling. Furthermore, process optimisation was performed using central composite design (CCD) and RSM, a method that has not been systematically applied to CPH-derived carbons. Beyond performance, this study highlights the green valorisation of an abundant agro-waste into chemically activated adsorbents, aligning with circular economy principles. By coupling enhanced graphitisation with optimised adsorption efficiency, the proposed approach highlights the green valorisation of an abundant agro-waste into chemically activated adsorbents, aligning a sustainable and cost-effective solution for wastewater remediation in cocoa-producing regions.
The raw material, referred to as CPH, comprised 4 kg of agricultural waste. The husks were thoroughly washed multiple times with deionized water to remove dust and surface impurities. Subsequently, the material was air-dried under sunlight for three days on a flat, open surface, followed by oven drying at 100 °C for 48 hours using a Riossa oven (Riossa, Mexico). After drying, the material was ground and sieved to obtain particles smaller than 0.25 mm (mesh size 60–80), using a conventional mechanical sieve shaker. Similar pre-treatment steps have been widely reported for lignocellulosic precursors to ensure homogeneity and surface cleanliness prior to activation.15,23,24 The resulting CPH powder was subjected to the determination of extractives, cellulose, hemicellulose, and lignin contents following the standard analytical protocols reported in the Official Methods of Analysis of AOAC International (18th ed.),42 in accordance with the acid–alkali digestion and solvent extraction procedures commonly employed for lignocellulosic residues [Sluiter et al., NREL/TP-510-42618, 2012].43 Comparable pre-treatment and characterisation methodologies for lignocellulosic precursors have also been reported by Selvaraju and Bakar (2017)44 and Li et al. (2014),45 whose studies are widely cited and validate the reliability of these analytical techniques.
:
1 v/v) and maintained at a constant temperature for 3 hours to extract soluble components. The mixture was then filtered, and the solid residue was dried in an oven at 105 °C for 24 hours until a constant weight was achieved. After cooling to room temperature in a desiccator, the sample was weighed (W1). The extractable content (% Ext) was calculated using eqn (1):
![]() | (1) |
The remaining residue was subsequently used for quantifying the hemicellulose and lignin contents.42–45
![]() | (2) |
![]() | (3) |
| % Cellulose = 100 − % Ext − % Hemi − % Lignina | (4) |
:
10 (w/v) ratio of lignocellulosic residue to chemical solution at room temperature for 12 hours, with two concentration levels applied for each condition. For alkaline activation, CPHP, previously oven-dried, was immersed in sodium hydroxide solutions at concentrations of 0.1 M and 5 M (solid-to-liquid ratio of 100 g L−1). The resulting samples were designated as ANCSH0.1 and ANCSH5, respectively. Similarly, acidic treatment was performed by immersing CPHP in sulphuric acid solutions at 0.5 M and 10 M concentrations, yielding samples labelled ANCSA0.5 and ANCSA10. After treatment, all suspensions were vacuum-filtered and thoroughly washed with deionised water until the filtrate reached a neutral pH, followed by rinsing with acetone. The treated solids were then oven-dried at 80 °C until constant weight was achieved, typically within 24 hours. Previous studies have demonstrated the effectiveness of alkali and acid pretreatments in modifying CPH for enhanced dye adsorption, particularly sodium hydroxide delignification to improve surface reactivity and porosity.20,23,24,33,34,36
00× to examine surface morphology and elemental distribution, which are critical parameters in evaluating biomass-derived carbons.46 The characterisation was conducted at the Center for Research in Science and Applied Technology of Tabasco (CICTAT), Academic Division of Basic Sciences, Universidad Juárez Autónoma de Tabasco (UJAT), in collaboration with the Faculty of Chemical Sciences at Benemérita Universidad Autónoma de Puebla (BUAP), with technical support from Dr Gerardo Enrique Córdova Pérez.
The average crystallite size was estimated using the Scherrer eqn (5).
![]() | (5) |
The adsorption capacity at equilibrium, qe (mg g−1), was calculated using eqn (6):
![]() | (6) |
The following expression was used to calculate the dye removal percentage.
![]() | (7) |
The pseudo-first-order kinetic model15 is expressed in linear form in eqn (8).
ln(qe − qt) = ln qe − k1t
| (8) |
And the nonlinear form is given by ref. 54:
| qt = qe(1 − e−k1t) | (9) |
The pseudo-second-order kinetic model60 in linear form is given by:
![]() | (10) |
The nonlinear form of the pseudo-second-order kinetic model61 is exposed in the following eqn (11).
![]() | (11) |
where the variables are the same as in the linear form.
These models assume different adsorption mechanisms: the pseudo-first-order model is typically associated with physisorption. In contrast, the pseudo-second-order model suggests chemisorption involving valence forces through sharing or exchange of electrons between the adsorbent and adsorbate.41,62 The data were processed using Origin 8.5 software. By analysing the kinetic parameters and coefficient of determination (R2), the best-fitting model can be determined, offering valuable insights into the controlling mechanism of the adsorption process.
The Langmuir isotherm64 assumes monolayer adsorption onto a surface with a finite number of identical sites and no interaction between adsorbed molecules. It is expressed as:
![]() | (12) |
The Freundlich isotherm54 is an empirical model that describes adsorption onto heterogeneous surfaces with a non-uniform distribution of heat of adsorption. Its eqn (13) is given by:
| qe = KFC1/ne | (13) |
The Redlich–Peterson model65 combines the features of both Langmuir and the Freundlich isotherms, offering flexibility for heterogeneous or hybrid adsorption systems. Its general form is:
![]() | (14) |
The Sips isotherm24 describes adsorption on heterogeneous surfaces and reduces to the Langmuir equation at low concentrations. Its eqn (15) is:
![]() | (15) |
The Temkin model63 accounts for adsorbate–adsorbent interactions and assumes the heat of adsorption decreases linearly with surface coverage. The eqn (16) is:
qe = B ln(ACe)
| (16) |
A is the Temkin isotherm equilibrium binding constant (L g−1),
, where bT is a constant related to the heat of adsorption (J mol−1), R is the universal gas constant (8.314 J mol−1 K−1), T is the absolute temperature (K).
The Toth isotherm66 is an empirical model used to describe adsorption on heterogeneous surfaces. It reduces to Langmuir when the heterogeneity parameter t = 1. The eqn (17) is:
![]() | (17) |
qmax is the maximum adsorption capacity (mg g−1), KT is the Toth equilibrium constant (L mg−1), t is the Toth heterogeneity parameter (dimensionless).
| α = 2k/4 | (18) |
| N = 2k + 2k + Cp | (19) |
![]() | (20) |
The CCD matrix for this study consisted of 12 experimental runs, including four center points to assess reproducibility and four axial points to ensure rotatability of the design (Table 1). The experimental runs were done at random to minimise systematic error. Analysis of variance (ANOVA) was used to evaluate the significance of individual and interaction effects, with main effects considered statistically significant at p < 0.05 (95% confidence level). All experiments were performed in duplicate, with each measurement conducted in triplicate to ensure accuracy and statistical reliability.
| Runs order | Temperature (°C) | pH |
|---|---|---|
| 1 | 35 | 2 |
| 2 | 35 | 6 |
| 3 | 65 | 2 |
| 4 | 65 | 6 |
| 5 | 29 | 4 |
| 6 | 71 | 4 |
| 7 | 50 | 1 |
| 8 | 50 | 7 |
| 9 | 50 | 4 |
| 10 | 50 | 4 |
| 11 | 50 | 4 |
| 12 | 50 | 4 |
Optimisation was carried out using the desirability function to identify the best pH and temperature conditions for maximum MB removal and adsorption capacity.
| Lignocellulosic composition | Content (wt%) | Proximate analysis | Content (wt%) |
|---|---|---|---|
| Extractable | 13.4 ± 0.47 | Humidity | 9.26 ± 0.02 |
| Hemicellulose | 47.73 ± 7.84 | Protein | 5.29 ± 0.17 |
| Lignin | 18.95 ± 9.61 | Fat | 7.31 ± 0.07 |
| Cellulose | 19.92 ± 1.39 | Ash (%) | 8.03 ± 0.10 |
| Fiber | 23.20 ± 1.36 | ||
| Carbohydrate | 46.9 ± 0.00 |
The chemical composition of lignocellulosic bioadsorbents significantly influences their adsorption performance, as it determines the availability of functional groups and surface chemistry. CPHP is primarily composed of natural fiber polymers, with hemicellulose being the most abundant carbohydrate fraction, followed by cellulose and lignin. The cellulose content measured (19.92 wt%) is lower than values reported in other studies,39,73 which may be attributed to differences in cocoa variety, cultivation practices, soil characteristics, and regional climatic conditions. In contrast, the hemicellulose content in this study was higher than that reported elsewhere, reinforcing the influence of agricultural and genetic factors on biomass composition.38,39
Additional components present in smaller quantities, such as proteins, ashes, and extractables, are grouped under “others” due to their secondary relevance in the structural composition of the biomass. Nevertheless, these constituents may still contribute to the surface reactivity of the material, particularly in chemical activation and adsorption processes.38
| Adsorbent | SBET (m2 g−1) | PV (cm3 g−1) | PS (nm) |
|---|---|---|---|
| CPH | 0.0169 | 0.000778 | 7 |
| ANCSA0.5 | 0.129 | 0.000366 | 10 |
| ANCSA10 | 1.966 | 0.004617 | 14 |
| ANCSH0.1 | 0.5115 | 0.000039 | 3.5 |
| ANCSH5 | 0.2218 | 0.000511 | N.D. |
![]() | ||
| Fig. 3 SEM micrographs of (a) CPH, (b) ANCSH0.1, (c) ANCSH5, and (d) ANCSA10. Experimental conditions: accelerating voltage 5–20 kV, magnifications 100×–1000×. | ||
The predominantly quasi-spherical and rod morphology of ANCSH5 and ANCSA10 particles is evident, with most seen particles exhibiting this shape. This morphological transformation is primarily attributed to the elimination of hydrophobic waxes, pectin, hemicellulose, lignin, and other impurities from the CPH matrix during sequential acid and alkaline treatments. The defibrillation of cellulose into individual fibrils is linked to the removal of its amorphous regions, predominantly influenced by the presence of hemicellulose and lignin.18,21,27 The total weight loss of the CPH, 18% after the NaOH treatment and 40% after the H2SO4 treatment, further confirms this.
Fig. 4 shows the EDX Spectra of CPH, ANCSH5, and ANCSA10. Elemental analysis of CPH reveals the presence of carbon (47.40%), oxygen (38.32%), nitrogen (13.93%), potassium (0.15%), phosphorus (0.04%), and calcium (0.16%). However, elemental analysis of ANCSH5 and ANCSA10 reveals the presence of all elements detected in CPH, except for phosphorus, potassium, and calcium. The increased carbon peaks (Fig. 4b and c) and higher estimated mass percentages (Table 4) seen in the EDX analysis of ANCSA10 and ANCSH5, relative to CPH, indicate that the chemical treatment effectively promoted the carbonization process.
| Chemical element (ms%) | CPH | ANCSA10 | ANCSH5 |
|---|---|---|---|
| C | 47.40 | 49.19 | 48.91 |
| N | 13.93 | 21.70 | 20.36 |
| O | 38.32 | 28.8 | 30.73 |
| K | 0.15 | 0.01 | N.D. |
| P | 0.04 | N.D. | N.D. |
| Ca | 0.16 | 0.30 | N.D. |
| Total | 100 | 100 | 100 |
Similar outcomes were obtained by Mousavi-Qeydary et al.75 in their study of activated carbon made from human hair waste, and Putra Negara et al.76 who synthesised activated carbons derived tabah bamboo.
Significant structural transformations were seen following chemical activation. The ANCSH5 sample (treated with 5 M NaOH) displayed sharper and more intense reflections at the same positions, indicating enhanced crystallinity and a higher degree of graphitization. Notably, the increased intensity of the (100) peak at 43.5° suggests the formation of graphene-like sp2 carbon domains, likely due to the effective delignification and removal of hemicellulose under strong alkaline hydrolysis.48
In contrast, the ANCSA10 sample (treated with 10 M H2SO4) showed additional peaks at 29.4° and 31.7°, alongside those at 21.7°, 34.4°, and 43.5°. These are attributed to the formation of organosulfate or sulfonated crystalline phases, reflecting more complex structural rearrangements. However, the broader peaks indicate that strong acid activation induced partial structural collapse, limiting long-range order.47,52
Samples treated with lower concentrations of acid (ANCSA0.5) and base (ANCSH0.1) exhibited only weak reflections at 34.4° and 43.5°, indicating limited carbon ordering and incomplete removal of amorphous constituents.
The prominent reflection at 43.5°, corresponding to the (100) plane of hexagonal graphite, serves as a key indicator of aromatic ring condensation and planar sp2 carbon alignment.48,78 Its pronounced intensity in ANCSH5 confirms the highest degree of graphitisation among all tested materials, consistent with its enhanced physicochemical properties for adsorption applications.
![]() | ||
| Fig. 6 Raman spectra of CPH, ANCSH0.1, ANCSH5, ANCSA0.5, and ANCSA10. Excitation wavelength: 632.8 nm. | ||
The presence of both D and G bands in all samples confirms the carbonaceous nature of the materials, with varying degrees of structural order. The D and G bands were seen in the ranges of 1300–1410 cm−1 and 1550–1610 cm−1, respectively, consistent with reported values in the literature for disordered carbon matrices.50,79,80 The D band, indicative of structural disorder, exhibited the highest intensity in the raw CPH sample, reflecting a highly disordered carbon framework. In contrast, this behaviour differed in the chemically treated samples, suggesting modifications in the carbon framework due to the treatments.81–83 The D band is associated with disordered carbon regions, and the G band corresponds to ordered graphitic structures. The D band arises from the breathing modes of sp2-hibridised carbon atoms in aromatic rings and is activated by structural defects or disorder, often associated with sp3 hybridisation. In contrast, the G band results from the in-plane stretching of sp2-hybridised carbon atoms, characteristic of ordered graphitic domains. These well-defined features in the Raman spectra offer valuable insight into the heterogeneous microstructure of the carbon materials.79,83,84
To evaluate the degree of graphitisation, the intensity ratio of the D peak and G peaks was calculated, where:
![]() | (21) |
In samples treated with sodium hydroxide, particularly ANCSH5, the D band intensity exceeded that of the G band, resulting in higher R values. In contrast, sulphuric acid-treated samples did not exhibit the same trend. Interestingly, higher R values, often associated with greater ID, may also reflect an increased presence of aromatic ring structures and a partial reduction in certain types of carbon-related defects.79,85,86 This structural evolution can enhance the formation of oxygen-containing functional groups on the material surface, thereby improving adsorption performance by increasing surface polarity and providing additional active sites for interaction with adsorbate molecules such as dyes. Overall, sodium hydroxide treatment promotes a higher degree of graphitisation compared to sulphuric acid treatment of the raw biomass.41,81
O stretching in ester or carboxylic groups, likely originating from hemicellulose or lignin.20,87 The signal at ∼1600 cm−1 is attributed to aromatic C
C stretching or conjugated C
O groups within lignin, and may also reflect π–π interactions associated with emerging graphitised domains.5,55,89 The peak at 1426 cm−1 corresponds to aromatic skeletal C
C stretching,90 while 1314 cm−1 is assigned to in-plane C
C vibrations and symmetric C–H bending of methyl/methylene groups, common in lignin side chains and polysaccharide residues.88 A strong band at 1244 cm−1 corresponds to C–O–C ether bonds in lignin,21 and the absorption at 1028 cm−1 is assigned to C–O stretching of primary and secondary alcohols in cellulose and hemicellulose.5,21 Finally, the band near 590 cm−1 corresponds to out-of-plane C–H bending vibrations.62
In contrast, the FTIR spectra of NaOH-treated samples (Fig. 7b and c) show a progressive attenuation and disappearance of the 1722 cm−1 band, particularly in the ANCSH5 sample, indicating cleavage of ester linkages and effective removal of hemicellulose.24 Similarly, the disappearance of bands at 1244 cm−1 and 590 cm−1 confirms lignin degradation. This depolymerization process exposes hydroxyl and phenolic groups, thereby enhancing surface reactivity.23 The alkaline treatment thus results in significant removal of recalcitrant fractions and increases cellulose accessibility, which is critical for adsorption efficiency.20 Notably, new bands appeared at 702 cm−1 (ANCSH0.1) and 704 cm−1 (ANCSH5), associated with out-of-plane C–H bending in aromatic rings, suggesting the development of π–π stacking interactions between carbon domains, possibly linked to partially graphitised structures.19,80
In the case of the acid-treated sample (Fig. 7d), the 1722 cm−1 band was red-shifted to 1704 cm−1, indicating possible alterations in carbonyl environments due to partial conjugation. While the 1244 cm−1 band (associated with either C–O-linkages in lignin) disappeared, suggesting partial lignin removal, the 590 cm−1 band remained, indicating incomplete depolymerization of aromatic structures. In addition, intensified bands in the 1210–1114 cm−1 region were seen, associated with C–OH and C–O–C stretching, which are indicative of newly formed oxygen-containing acidic functionalities such as lactones, anhydrides, and phenols.24,55,89 Notably, unlike the alkali-treated samples (ANCSH0.1 and ANCSH5), which exhibited new bands at 702 and 704 cm−1, respectively, attributed to π–π stacking interactions, these bands were absent in the acid-treated ANCSA10. The disappearance of the 1314 cm−1 band in ANCSA10 further supports the conclusion that sulphuric acid treatment was less effective for delignification and the development of aromatic carbon structures.
The band at 1028 cm−1, attributed to C–O stretching vibrations of primary and secondary alcohols in cellulose and hemicellulose, remained prominent after sulphuric acid treatment (ANCSA10), but showed a notable increase in intensity in the NaOH-treated samples (ANCSH0.1 and ANCSH5). This enhancement suggests a more effective removal of hemicellulose and lignin by the alkaline treatment, resulting in greater exposure of cellulose hydroxyl groups. These newly accessible polar functionalities can contribute significantly to adsorption performance by enabling electrostatic interactions and hydrogen bonding with cationic pollutants. Similar trends have been reported in other studies, where alkaline pretreatment led to a stronger 1028 cm−1 signal due to increased cellulose accessibility.59,91,92
The more effective delignification achieved through NaOH treatment is attributed to enhanced cellulose exposure and the formation of surface oxygen-containing functional groups such as carbonyl, carboxyl, and hydroxyl groups. These functionalities are known to enhance adsorption performance through mechanisms such as electrostatic attraction, hydrogen bonding, and metal–ligand complexation.5 According to previous studies, the nature and abundance of surface functional groups significantly influence the adsorption mechanism, with –COOH and –OH groups promoting the uptake of polar and ionic contaminants, including dyes and heavy metals.55,89 Therefore, the applied chemical treatments effectively altered the surface chemistry of CPH, promoting the generation of active sites and improving the material's affinity toward the target contaminant.
Although pure graphitic carbon exhibits minimal infrared activity due to its symmetrical and non-polar structure, partially graphitised domains formed during thermal activation of cellulose produce discernible FTIR bands. These include aromatic C
C stretching near 1600–1610 cm−1, for all materials, as well as bands near 1415–1426, 1314–1320, for CPH, ANCSH0.1, and ANCSSH5; and 702–704 cm−1. These latter bands are exclusively presented for ANCSH0.1 and ANCSH5, which are associated with sp2-hybridised carbon structures and π–π interactions, evidencing the alkaline treatment effect. These features reflect the evolution of oxygenated amorphous carbon networks derived from lignocellulosic precursors. Notably, lignin's inherent aromatic ring structures contribute to π-conjugated domains, facilitating π–π stacking interactions with aromatic pollutants such as cationic methylene blue (MB+), which enhances adsorption efficiency, especially in alkaline-treated materials.24,40,41,89
FTIR analysis corroborated the structural changes identified by Raman spectroscopy and XRD. The presence of bands near 1600 cm−1 (C
C stretching) and 1220–1260 cm−1 and 1028 cm−1 (C–O stretching) confirms the formation of conjugated aromatic systems and oxygen-rich functionalities during chemical activation. These observations align with the D and G bands in Raman spectra, which signify the coexistence of disordered and graphitic carbon domains. Additionally, the broad peaks in the XRD patterns indicate the presence of amorphous or turbostratic carbon, supporting the formation of partially graphitised structures. Together, these spectroscopic and diffraction analyses demonstrate the successful conversion of CPH into a carbonaceous adsorbent with structural features favorable for π–π interactions and electrostatic adsorption mechanisms.5,20,41,55
Beyond adsorption efficiency, the sustainability of ANCSH5 is intrinsically linked to the valorisation of CPH, a locally abundant agro-waste with limited alternative applications. Its conversion into carbon prevents biomass accumulation in landfills, where uncontrolled biological decomposition generates methane (CH4), a potent greenhouse gas.22,38 In addition, the synthesis employed here requires only sulphuric acid and sodium hydroxide pretreatments at moderate conditions, without the need for high-temperature furnaces, thus minimizing energy consumption and the associated carbon footprint.39,47 These attributes distinguish the present route from other biomass-based carbons and further from petroleum-derived ACs, whose production pathways involve fossil feedstocks with significant emissions of GreenHouse Gases (GHGs) and Volatile Organic Compounds (VOCs).15,77 This highlights the “green” positioning of the proposed synthesis, which integrates waste reduction, low-energy processing, and circular economy principles.
This improvement may be attributed to the removal of organic compounds during the delignification process facilitated by alkaline treatment, which enhances the chelating ability of the adsorbent surface. Additionally, sodium hydroxide activation promotes a higher degree of graphitization compared to sulphuric acid treatment, leading to the formation of oxygen-containing functional groups. These modifications enhance surface polarity and create additional active sites, thereby enhancing the overall adsorption performance.
![]() | ||
| Fig. 8 Adsorption kinetics of MB on (a) CPH, (b) ANCSA10, and (c) ANCSH5. Experimental conditions: C0 = 5 mg L−1, pH 6, T = 30 °C. | ||
| Adsorbent | Pseudo-first-order kinetics | ||
|---|---|---|---|
| k1 (min−1) | qe (mg g−1) | R2 | |
| CPH | 0.02513 | 0.21966 | 0.99826 |
| ANCSH5 | 0.03045 | 0.71546 | 0.96173 |
| ANCSA10 | 0.02825 | 0.99607 | 0.65892 |
| Adsorbent | Pseudo-second-order kinetics | ||
|---|---|---|---|
| k2 (g mg−1 min−1) | qe (mg g−1) | R2 | |
| CPH | 0.12188 | 0.27490 | 0.99827 |
| ANCSH5 | 0.11569 | 1.18085 | 0.99918 |
| ANCSA10 | 0.00005 | 7.57070 | 0.98107 |
| Adsorbent | Langmuir isotherm | ||
|---|---|---|---|
| qmax (mg g−1) | KL (L mg−1) | R2 | |
| CPH | 100.0000 | 0.002418 | 0.5291 |
| ANCSH5 | 100.0000 | 0.024424 | 0.9997 |
| ANCSA10 | 100.0000 | 0.007475 | 0.1416 |
| Adsorbent | Freundlich isotherm | ||
|---|---|---|---|
| KF (mg1−1/n g−1 L1/n) | n | R2 | |
| CPH | 0.000016 | 0.201033 | 0.8449 |
| ANCSH5 | 2.377979 | 1.037309 | 0.9999 |
| ANCSA10 | 0.356150 | 0.687208 | 0.1559 |
The Langmuir model, which assumes monolayer adsorption onto a homogeneous surface with finite, identical sites, showed a strong correlation for ANCSH5 (R2 = 0.9997). This suggests that the adsorption onto ANCSH5 may involve more uniform active sites with higher energy affinity toward MB. In contrast, CPH and ANCSA10 yielded low Langmuir R2 values (0.53 and 0.14, respectively), indicating that the assumption of surface homogeneity does not hold for these materials.
The Freundlich model, which accounts for adsorption on heterogeneous surfaces, provided a better fit for CPH (R2 = 0.85) and ANCSH5 (R2 = 0.9999), while it still performed poorly for ANCSA10 (R2 = 0.16). These differences may be attributed to variations in surface functionalization and porosity induced by chemical treatment, as supported by FTIR and BET analyses.
As shown in Fig. 9, the experimental adsorption data were fitted using the non-linear forms of the Langmuir and the Freundlich isotherm models. The results indicate an excellent fit for ANCSH5, moderate agreement for CPH, and poor conformity for ANCSA10, highlighting differences in surface characteristics and adsorption mechanisms among the adsorbents. In particular, CPH and ANCSA10 exhibited non-ideal and progressive adsorption behaviour. These deviations can be attributed to surface heterogeneity, low adsorbate–adsorbent affinity, and multilayer adsorption factors not fully captured by the monolayer assumption of the Langmuir model or the empirical approximation inherent to the Freundlich model. Importantly, although the equilibrium data showed deviations from classical isotherms, the adsorption process itself was effective, as evidenced by the high MB removal percentages and the excellent agreement of the kinetic data with the pseudo-second-order model. This reinforces the idea that adsorption performance should not be evaluated solely based on isotherm conformity, but also considering the dynamic interaction mechanisms and material-specific properties.
![]() | ||
| Fig. 9 Experimental adsorption isotherms (qe vs. Ce) of MB on (a) CPH, (b) ANCSA10, and (c) ANCSH5, fitted with Langmuir and Freundlich models. | ||
To further explore the equilibrium behaviour of MB adsorption onto the three adsorbents, extended isotherm models, including the Sips, Redlich–Peterson, Temkin, and Toth models, were applied. These models are useful for describing adsorption on heterogeneous surfaces, mixed adsorption mechanisms, or multilayer phenomena not captured by classical models. Fig. 10 shows the plots for the extended isotherm models (Redlich–Peterson, Sips, Temkin, and Toth), and the calculated isotherm constants are summarised in Table 7.
![]() | ||
| Fig. 10 Experimental adsorption isotherms (qe vs. Ce) of MB on (a) CPH, (b) ANCSH5, and (c) ANCSA10, fitted with extended isotherm models (Redlich–Peterson, Sips, Temkin, and Toth). | ||
| Adsorbent | Sips isotherm | |||
|---|---|---|---|---|
| qmax (mg g−1) | Ks (L mg−1) | n | R2 | |
| CPH | 100.0000 | 0.0443 | 5.0955 | 0.8444 |
| ANCSH5 | 100.0000 | 0.0248 | 1.0053 | 0.9997 |
| ANCSA10 | 3.9212 | 0.2869 | 10.0000 | 0.2948 |
| Adsorbent | Redlich–Peterson isotherm | |||
|---|---|---|---|---|
| KR (L g−1) | aR | g | R2 | |
| CPH | 18.0386 | 75.3073 | 0.0000 | 0.5357 |
| ANCSH5 | 100.0000 | 41.0527 | 0.0368 | 0.9999 |
| ANCSA10 | 42.2199 | 57.4525 | 0.0000 | 0.1431 |
| Adsorbent | Toth isotherm | |||
|---|---|---|---|---|
| qmax (mg g−1) | KT (L mg−1) | t | R2 | |
| CPH | 2.0345 | 14.4479 | 23.4162 | 0.0000 |
| ANCSH5 | 100.0000 | 0.0241 | 1.0793 | 0.9998 |
| ANCSA10 | 13.7814 | 0.0524 | 60.5573 | 0.1431 |
| Adsorbent | Temkin isotherm | ||
|---|---|---|---|
| A (L g−1) | B | R2 | |
| CPH | 0.2207 | 3.0693 | 0.6126 |
| ANCSH5 | 2.6914 | 2.9228 | 0.9306 |
| ANCSA10 | 0.4074 | 5.3564 | 0.1959 |
The Sips model, which combines elements of Langmuir and Freundlich, yielded excellent correlation for ANCSH5 (R2 = 0.9997), consistent with its previously seen near-monolayer behaviour with some degree of heterogeneity. For CPH, the model improved the fit moderately (R2 = 0.84), suggesting a surface with variable site energies. In contrast, ANCSA10 showed a limited correlation (R2 = 0.29), indicating that the model assumptions may not apply to its surface characteristics.
The Redlich–Peterson model, capable of bridging both Langmuir and Freundlich forms, also provided an exceptional fit for ANCSH5 (R2 = 0.9999). Its performance was slightly lower for CPH and poor for ANCSA10, again reflecting heterogeneity and non-ideal adsorption patterns. The Temkin model, which assumes a linear decrease in adsorption energy with coverage, showed moderate fits for CPH and ANCSH5, while underperforming for ANCSA10. This suggests that adsorbent–adsorbate interactions in these systems involve energetic heterogeneity or weak bonding effects. The Toth model, often used for highly heterogeneous systems, resulted in the best fit for ANCSH5 (R2 ≈ 1) and improved fit for CPH compared to classical models. However, ANCSA10 again showed poor adjustment, highlighting its distinct and possibly multilayer or diffusion-limited adsorption mechanism.
ANCSH5 exhibited strong agreement with all models, particularly Redlich–Peterson and Toth, indicating a highly efficient and energetically accessible surface. The similarity in performance among classical and extended models for ANCSH5 suggests well-distributed active sites, moderate surface heterogeneity, and favorable interaction with the adsorbate. In contrast, ANCSA10 showed poor correlation across all models, indicating deviation from ideal or energetically uniform adsorption behaviour likely due to severe surface heterogeneity, inaccessible sites, or multilayer interactions. CPH showed moderate fits, with Sips and Temkin performing better.
In summary, the application of extended isotherm models supports the conclusion that ANCSH5 possesses a surface with high affinity and partially homogeneous active sites, whereas CPH shows heterogeneous behaviour with moderate adsorption energy. The poor fit of all models to ANCSA10 underscores the need for either customised models or hybrid approaches for complex surfaces that incorporate kinetic or diffusional constraints.
| Adsorbent | Langmuir | Freundlich | Sips | Redlich–Peterson | Temkin | Toth |
|---|---|---|---|---|---|---|
| CPH | 0.5291 | 0.8449 | 0.8444 | 0.5357 | 0.6126 | 0.0000 |
| ANCSA10 | 0.1416 | 0.1559 | 0.2948 | 0.1431 | 0.1959 | 0.1431 |
| ANCSH5 | 0.9997 | 0.9999 | 0.9997 | 0.9999 | 0.9306 | 0.9998 |
![]() | ||
| Fig. 11 Comparison of R2 obtained from classical and extended isotherm models for MB adsorption onto CPH, ANCSA10, and ANCSH5. | ||
Among the three adsorbents, ANCSH5 consistently exhibited the highest R2 values across all models, with nearly perfect fits for the Sips, Toth, and Redlich–Peterson models (R2 > 0.999). This indicates a well-defined and partially homogeneous adsorption mechanism, likely involving both high surface area and active functional groups enhanced by NaOH activation.
In contrast, CPH showed moderate fits, with the Freundlich and Sips models providing the best results (R2 ≈ 0.84). The relatively low R2 values for Langmuir and Redlich–Peterson suggest that adsorption onto CPH occurs on a heterogeneous surface without saturation, likely due to limited availability or diversity of active sites.
ANCSA10 exhibited the lowest R2 values across nearly all models, including extended ones. This suggests a highly complex or non-ideal adsorption process, potentially involving multilayer formation, diffusion limitations, or weak interactions.
These findings reinforce the importance of applying a range of isotherm models, classical and extended, when characterizing novel adsorbents. Relying solely on Langmuir or Freundlich models may obscure underlying mechanistic differences, particularly in bio-based or structurally diverse materials, and reveal significant differences in adsorption capacity and surface affinity among the adsorbent materials.
ΔG0 = −RT ln KC
| (22) |
![]() | (23) |
| ΔG0 = ΔH0 − TΔS0 | (24) |
Table 9 summarises the calculated thermodynamic parameters. Negative values of ΔG° across all tested temperatures confirm the spontaneous nature of MB adsorption onto ANCSH5, consistent with observations reported for other biomass-derived adsorbents.9,57,95 The positive ΔH° (25.0 kJ mol−1) indicates an endothermic process, while positive ΔS° values reflect increased randomness at the solid–liquid interface, in agreement with previous studies on dye adsorption involving natural adsorbents.64 The magnitude of ΔH° suggests that the adsorption involves strong physicochemical interactions, consistent with electrostatic attraction, hydrogen bonding, and π–π stacking mechanisms previously identified. These results corroborate the kinetic and isotherm analyses, demonstrating that MB adsorption on ANCSH5 is thermodynamically favorable, spontaneous, and enhanced at higher temperatures.9,57,95
| T (°C) | T (K) | ΔG° (kJ mol−1) | ΔH° (kJ mol−1) | ΔS° (J mol−1 K−1) |
|---|---|---|---|---|
| 25 | 298.15 | −4.76 | 25.0 | 99.8 |
| 35 | 308.15 | −5.52 | 25.0 | 99.6 |
| 45 | 318.15 | −6.28 | 25.0 | 99.1 |
All experiments were conducted at pH 6, a condition near the slightly acidic regime, which can influence surface charge and adsorbate speciation. At this pH, functional groups such as hydroxyls and carboxyls on the adsorbent surface may be partially deprotonated, enhancing electrostatic attraction with cationic species or promoting hydrogen bonding. This effect is particularly favorable in materials with higher surface polarity and accessible acidic groups.
Among all tested adsorbents, ANCSH5 demonstrated the highest performance, with excellent agreement across all isotherm models (R2 > 0.99). The strong affinity seen at pH 6 indicates a surface rich in accessible functional groups whose ionization state promotes adsorption through electrostatic interactions and hydrogen bonding. These interactions are likely supported by the mildly acidic conditions, which optimise the charge distribution and polarity of the adsorbate–adsorbent system. This behaviour aligns with reports on functionalised bio-carbons, where pH-induced deprotonation increases the uptake of cationic dyes.96
CPH, while exhibiting moderate adsorption capacity, was best described by the Freundlich model (R2 = 0.788), suggesting adsorption on a heterogeneous surface with sites of varying energy. The lower R2 values obtained for the Langmuir and extended models indicate that adsorption at pH 6 deviates from ideal monolayer behaviour, potentially due to partial ionization of surface groups or weak adsorbate–adsorbent interactions under mildly acidic conditions. This limited affinity aligns with prior studies on lignocellulosic adsorbents in similar pH environments.97,98
In contrast, ANCSA10 showed poor correlation with all isotherm models, including Sips and Toth (R2 < 0.4), indicating low adsorption capacity and inadequate model description at pH 6. This behaviour may result from surface deactivation, a low density of functional groups, or steric hindrance. The lack of improvement under acidic conditions suggests that its active sites are unresponsive to moderate pH shifts or that other mechanisms, such as deep microporosity, intraparticle diffusion, or non-electrostatic interactions, may dominate. Such deviations from equilibrium-based models have been previously reported for bio-based adsorbents with complex surface architectures.99
These findings highlight the crucial role of solution pH in regulating adsorption efficiency, particularly for bio-based adsorbents with heterogeneous surface chemistries. Under uniform pH 6 conditions, ANCSH5 exhibited superior performance, attributed to its surface functionality being well-aligned with mildly acidic environments. In contrast, CPH displayed a moderate adsorption response, while ANCSA10 remained largely unresponsive.
The acid dissociation constant (pKa) of MB is approximately 3.8, indicating that above this pH, the dye predominantly exists in its cationic form (MB+) in aqueous solution.100–102 Consequently, under the experimental conditions employed in this work (pH 6–7), MB molecules are positively charged and readily interact with negatively charged adsorbent surfaces. When the solution pH is equal to or higher than the pHPZC (6.2 for ANCSH5 and 4.8 for ANCSA10), deprotonation of surface –COOH and –OH groups (–COOH → –COO−; –OH → –O−) generates a net negative charge, promoting strong electrostatic attraction with MB+, in addition to hydrogen bonding and π–π stacking with aromatic domains on the carbon surface.
At near-neutral pH, ANCSH5 therefore exhibits superior adsorption capacity, consistent with its higher pHPZC and greater availability of deprotonated oxygenated groups. Within the working pH window (6–7), the surface of ANCSH5 is partially deprotonated, exposing –COO− and –O− sites that enhance electrostatic interactions with MB+, synergistically reinforced by π–π stacking between the dye's aromatic rings and the partially graphitised domains of the adsorbent.38,52,55 In contrast, ANCSA10, with a lower pHPZC, remains closer to a neutral or slightly positive surface state, limiting the availability of negatively charged sites and thus reducing electrostatic affinity for MB+.12,26,51,54 This interplay between the dye's pKa and the adsorbent's pHPZC is a decisive parameter governing the charge-controlled adsorption mechanism and explains the high adsorption efficiency observed at near-neutral pH.
Table 10 compares the maximum adsorption capacities of ACs derived from various lignocellulosic biomasses for MB removal, predicted by isotherm modelling. Although some biomass-based carbons report higher qmax values (e.g., coffee husk, pineapple residues), these capacities were typically obtained under extreme conditions, such as very high initial MB concentrations (≥400 mg L−1) or excessive adsorbent dosages. By contrast, the NaOH-activated CPH (ANCSH5) synthesised in this work achieved a competitive qmax of 100 mg g−1 under more realistic operating conditions (C0 = 5–30 mg L−1, pH 6, 35 °C, dosage = 4 g L−1). This result is particularly relevant for practical wastewater treatment, where pollutant concentrations are moderate and the use of large adsorbent doses is economically unfeasible. Moreover, the integration of statistical optimisation via CCD-RSM adds a further competitive edge by ensuring that adsorption performance is guided by validated optimal conditions. The distinctive advantage of ANCSH5 compared with other biomass-derived carbons lies in the unique lignocellulosic composition of CPH, which is notably rich in hemicellulose. This facilitates extensive delignification during alkaline treatment and promotes the development of partially graphitised domains and a high density of oxygen-containing functional groups, as confirmed by Raman, XRD, and FTIR analyses. Unlike conventional precursors such as rice husk or coconut shell, where performance is largely governed by surface area, CPH-derived ANC combines enhanced graphitization with tailored surface chemistry. This synergy provides abundant active sites for electrostatic interactions, hydrogen bonding, and π–π stacking with MB molecules. Collectively, these features distinguish ANCSH5 as a competitive, sustainable, and industrially relevant adsorbent within the broader family of biomass-based carbons.
| Biomass precursor | Activation/treatment | qmax (mg g−1) | Conditions | Ref. |
|---|---|---|---|---|
| a MB C0: MB initial concentration range. | ||||
| Coffe husk | Pyrolysis | 416.68 | MB C0 range: 50–500 mg L−1; pH: 7; T: 30 °C; adsorbent dosage: 1 g L−1 | 5 |
| Pineapple leaf powder | Untreated | 281 | MB C0 range: 1.6–6.40 mg L−1; pH: 7.5; T: 24 °C; adsorbent dosage: 0.3 g L−1 | 51 |
| Pineapple waste (leaves, stem, crown) | Pyrolysis | 288.34 | MB C0 range: 5–400 mg L−1; pH: 7; T: 30 °C; adsorbent dosage: 10 g L−1 | 11 |
| Watermelon rings | H2SO4 | 200 | MB C0 range: 50–400 mg L−1; pH: 5.6; T: 30 °C; adsorbent dosage: 0.8 g L−1 | 55 |
| Sugarcane bagasse | Pyrolysis | 113.01 | MB C0 range: 20–50 mg L−1; pH: 7.4; T: 30 °C; adsorbent dosage: 0.6 g L−1 | 3 |
| Banana stem | Pyrolysis | 101 | MB C0 range: 25–200 mg L−1; pH: 7; T: 30 °C; adsorbent dosage: 3 g L−1 | 54 |
| Cocoa pod husk | NaOH | 100 | MB C0 range: 5–30 mg L−1; pH: 6; T: 35 °C; adsorbent dosage: 4 g L−1 | This study |
| Yellow passion-fruit peel | Untreated | 44.7 | MB C0 range: 5–600 mg L−1; pH: 8; T: 25 °C; adsorbent dosage: 10 g L−1 | 57 |
| Annona squmosa seed | H2SO4 | 25.91 | MB C0 range: 25–200 mg L−1; pH: 6; T: 27 °C; adsorbent dosage: 4 g L−1 | 96 |
| Brazil nut shell | Untreated | 7.81 | MB C0 range: 100–1500 mg L−1; pH: 6.5; T: 30 °C; adsorbent dosage: 100 g L−1 | 99 |
O) moieties, which are commonly formed through oxidation reactions during treatment with either acidic or alkaline agents. At pH 6, carboxylic groups tend to deprotonate, resulting in negatively charged carboxylate (–COO−) species, whereas hydroxyl and carbonyl groups contribute to surface polarity and hydrogen bonding capacity. Together, these functionalities enhance the overall hydrophilicity and facilitate electrostatic attraction with the MB+ molecules. This explains the improved adsorption performance seen in the NaOH-treated ANCSH5, in contrast to the H2SO4-activated ANCSA10, which likely retains a more protonated and less negatively charged surface under the same conditions.FTIR analysis revealed that the band at 1028 cm−1, assigned to the C–O stretching of primary and secondary alcohols, became more intense following NaOH treatment, indicating a higher density of exposed hydroxyl groups on the surface. These –OH groups, primarily from cellulose, play a critical role in the adsorption mechanism by participating in hydrogen bonding and electrostatic interactions with the MB+ molecules. This is particularly relevant at pH 6, where partial deprotonation of hydroxyl and carboxyl groups enhances surface polarity and negative charge. The increased exposure of C–OH groups, especially in ANCSH5, correlates with the improved adsorption performance seen in kinetic and isotherm studies. Similar findings have been reported in the literature, where stronger C–O bands at ∼1028 cm−1 were associated with improved adsorption properties due to enhanced accessibility of cellulose chains after alkaline pretreatment.59,91,92
At pH 6, π–π stacking interactions are particularly relevant for the adsorption of MB+, given its planar aromatic structure.24,41,55,89 The FTIR spectra revealed stronger aromatic C
C bands for ANCSH5, indicating a higher density of conjugated π-systems capable of forming stable interactions with MB+. These π–π interactions provide an additional adsorption mechanism beyond electrostatic attraction, especially in cases where the surface charge is neutral or weakly positive (as in ANCSA10). In ANCSH5, both π–π stacking and electrostatic forces act synergistically, leading to enhanced adsorption capacity and model fitting. Furthermore, this pH was selected to simulate mildly acidic wastewater conditions and to promote adsorption via electrostatic attraction, hydrogen bonding, and π–π stacking.
Fig. 13 illustrates three primary interactions:
![]() | ||
| Fig. 13 Proposed adsorption mechanism of MB+ onto oxygen-functionalised carbon-based adsorbents at pH 6. | ||
(1) π–π stacking, involving the aromatic rings of MB+ and conjugated π-domains on the adsorbent surface.
(2) Electrostatic attraction, driven by the deprotonation of carboxylic groups (–COOH → –COO−) on the adsorbent at pH 6, which interacts with the positively charged MB+ molecules.
(3) Hydrogen bonding, between surface hydroxyl (–OH) or carbonyl (–C
O) groups and electronegative atoms of MB+.
These combined interactions explain the strong affinity of MB+ toward materials like CPH and ANCSH5, where both surface polarity and negative charge enhance adsorption efficiency.
| % Removal = 9.254 + 2.203T − 1.596 pH + 0.191T pH − 0.0338T2 − 0.2052 pH2 | (25) |
| % q = 0.0900 + 0.0283T + 0.0239 pH + 0.0023T pH−0.0004T2 − 0.0028 pH2 | (26) |
| Parameters | Response | |||||
|---|---|---|---|---|---|---|
| (% Removal, Y1) | (Adsorption capacity (q), mg g−1, Y2) | |||||
| Runs order | Temperature, X1 (°C) | pH, X2 | Experimental | Predicted | Experimental | Predicted |
| 1 | 35 | 2 | 59.51 | 60.70 | 0.74 | 0.76 |
| 2 | 35 | 6 | 90.70 | 87.23 | 1.13 | 1.09 |
| 3 | 65 | 2 | 35.83 | 36.88 | 0.45 | 0.46 |
| 4 | 65 | 6 | 89.92 | 86.31 | 1.12 | 1.07 |
| 5 | 29 | 4 | 68.82 | 69.96 | 0.86 | 0.87 |
| 6 | 71 | 4 | 51.32 | 52.65 | 0.64 | 0.66 |
| 7 | 50 | 1 | 47.90 | 45.87 | 0.60 | 0.57 |
| 8 | 50 | 7 | 98.66 | 100 | 1.23 | 1.28 |
| 9 | 50 | 4 | 75.13 | 76.20 | 0.94 | 0.95 |
| 10 | 50 | 4 | 78.12 | 76.20 | 0.98 | 0.95 |
| 11 | 50 | 4 | 76.14 | 76.20 | 0.95 | 0.95 |
| 12 | 50 | 4 | 75.20 | 76.20 | 0.94 | 0.95 |
Tables 12 and 13 summarise the ANOVA results for the quadratic response surface models. The ANOVA for the sorption of MB onto ANCSH5 confirmed the overall statistical significance of the models, with F-values of 20.36 for % Removal and 20.69 for adsorption capacity (q, mg g−1). The corresponding p-values (<0.0001) indicate that both regression models are highly significant. Moreover, the R2, adjusted R2, and predicted R2 values for both models exceeded 0.97, demonstrating excellent model fit, generalizability, and predictive reliability.
| Source | Sum of squares | df | Mean square | F-value | p-value |
|---|---|---|---|---|---|
| a R2 = 0.9852, adjusted R2 = 0.9728, predicted R2 = 0.9852; p < 0.05 indicates the model terms are significant; P < 0.01 indicates the model terms are highly significant. | |||||
| Model | 985.48 | 5 | 197.10 | 20.36 | 0.0011 |
| X1-temperature | 268.62 | 1 | 268.62 | 27.75 | 0.0019 |
| X2-pH | 226.05 | 1 | 226.05 | 23.35 | 0.0029 |
| X1X2 | 131.10 | 1 | 131.10 | 13.54 | 0.0103 |
| X12 | 354.52 | 1 | 354.52 | 36.62 | 0.0009 |
| X22 | 5.19 | 1 | 5.19 | 0.54 | 0.4916 |
| Residual | 58.08 | 6 | 9.68 | ||
| Source | Sum of squares | df | Mean square | F-value | p-value |
|---|---|---|---|---|---|
| a R2 = 0.9850, adjusted R2 = 0.9725, predicted R2 = 0.9850; p < 0.05 indicates the model terms are significant; p < 0.01 indicates the model terms are highly significant. | |||||
| Model | 0.16 | 5 | 0.03 | 20.69 | 0.0010 |
| X1-temperature | 0.04 | 1 | 0.04 | 28.57 | 0.0018 |
| X2-pH | 0.04 | 1 | 0.04 | 23.78 | 0.0028 |
| X1X2 | 0.02 | 1 | 0.02 | 12.95 | 0.0114 |
| X12 | 0.06 | 1 | 0.06 | 37.51 | 0.0009 |
| X22 | 0.00 | 1 | 0.00 | 0.65 | 0.4512 |
| Residual | 0.01 | 6 | 0.00 | ||
For % removal, the statistically significant model terms were X1 (temperature), X2 (pH), X12 (temperature2), and the interaction term X1X2 (temperature × pH). Both the linear and quadratic terms of temperature were highly significant (p < 0.001), underscoring their critical influence on removal efficiency. The linear term for pH was also statistically significant (p < 0.01), whereas its quadratic term was not (p > 0.4). This suggests that pH affects the system primarily through linear behaviour and synergistic interaction with temperature, rather than exhibiting a pronounced curvilinear effect. The interaction term was statistically significant (p < 0.05), confirming a synergistic influence between temperature and pH.
For q (mg g−1), a similar trend was seen. The significant terms were X1, X12, and X1X2. Temperature remained the dominant factor, with both linear and quadratic terms showing strong statistical significance (p < 0.001). The interaction term was also significant (p < 0.05), whereas neither the linear nor the quadratic term of pH was. These findings reinforce the conclusion that adsorption capacity is primarily governed by temperature, with pH exerting an indirect effect through interaction. The inclusion of interaction and quadratic terms is crucial for capturing surface curvature and enhancing model accuracy, particularly for complex adsorption behaviors.
In summary, temperature emerged as the most influential variable in both models, while pH played a secondary but meaningful role through its interaction with temperature. These results validate the use of quadratic models within RSM for process optimisation and provide a robust statistical foundation for predictive modelling in dye adsorption systems. The significant F- and R2-values further support the inclusion of interaction and quadratic terms to capture the system's complexity.
Both response variables, % removal and adsorption capacity (q), were strongly affected by temperature and its nonlinear behaviour. The temperature × pH interaction contributed to refining the curvature of the response surface. Although pH alone was not statistically significant in the q model, and only marginally significant in the % removal model, its inclusion improved the overall model fit and predictive performance.
To evaluate the adequacy of the fitted quadratic models, residual diagnostics were conducted for both % removal and q. The assessment included Q–Q plots, histograms of residuals, residuals versus fitted values, and normality testing using the Shapiro–Wilk test.
For the % removal model (Fig. 14), the Q–Q plot showed that the residuals closely followed the theoretical normal distribution. The histogram of residuals appeared symmetric, and the residuals versus fitted values displayed random scatter without discernible patterns, supporting the assumption of homoscedasticity. The Shapiro–Wilk test returned a p-value of 0.1224, exceeding the 0.05 threshold, thus confirming the normality of residuals.
![]() | ||
| Fig. 14 Diagnostic plots for the quadratic model of % removal: Q–Q plot, histogram of residuals, residuals vs. fitted values, and Shapiro–Wilk normality test (p = 0.1224). | ||
Similarly, the diagnostic plots for the q model (Fig. 15) were consistent with the model assumptions. Residuals were normally distributed and randomly dispersed around zero across the range of fitted values. The Shapiro–Wilk test yielded a p-value of 0.1183, again supporting the normality assumption.
![]() | ||
| Fig. 15 Diagnostic plots for the quadratic model of q: Q–Q plot, histogram of residuals, residuals vs. fitted values, and Shapiro–Wilk normality test (p = 0.1224). | ||
Collectively, these diagnostic results confirm that both models satisfy the key assumptions of regression analysis: normality, homoscedasticity, and independence of residuals. Meeting these criteria ensures the validity of hypothesis testing, confidence intervals, and optimisation results derived from the models. Accordingly, the response surfaces for both % removal and q can be interpreted with high statistical confidence.
Optimisation of the response surface models indicated that both % removal and q achieve their maximum predicted values under comparable operational conditions. For % removal, the optimal conditions were 49.6 °C and pH 6.0, yielding a predicted removal efficiency of 99.05%. Similarly, for q, the optimum was also found at 49.6 °C and pH 6.0, with a corresponding predicted adsorption capacity of 1.24 mg g−1.
The three-dimensional response surface plots, along with their corresponding contour diagrams, provide detailed visual insight into the interactive effects of temperature and pH on MB removal efficiency (% removal) and adsorption capacity (q, mg g−1). These graphical representations illustrate the curvature of the response surfaces and facilitate the identification of optimal operating conditions for both responses. The seen surface curvature and contour gradients emphasise the synergistic influence of temperature and pH, reinforcing their combined impact on system performance (Fig. 16 and 17).
![]() | ||
| Fig. 16 Response surface 3D (a) and contour 2D (b) plots showing the interaction of temperature and pH on % MB removal. | ||
![]() | ||
| Fig. 17 Response surface 3D (a) and contour 2D (b) plots showing the interaction of temperature and pH on q. | ||
In the response surface for % removal (Fig. 16), a pronounced curvature is seen, particularly along the temperature axis, confirming the strong nonlinear effect of temperature. The response increases with temperature up to an optimal region (∼49.6 °C), beyond which it gradually plateaus, likely due to the saturation of available adsorption sites. The influence of pH is more subtle but plays a role in fine-tuning the response, primarily through its interaction with temperature.
The corresponding contour plot for % removal reinforces these observations by delineating zones of high efficiency in the central region of the design space. A region of maximum removal efficiency (>95%) is centered around moderate to high temperatures (approximately 48–55 °C) and slightly acidic to neutral pH values (5.5–6.5). The presence of elongated, diagonally oriented elliptical contour lines highlights a significant interaction between temperature and pH, where small changes in pH at a given temperature can markedly affect the removal efficiency.
Similarly, the response surface plot for q exhibits a curved topography (Fig. 17), with a distinct maximum occurring at intermediate levels of both temperature and pH. Temperature again emerges as the dominant factor, while pH contributes to enhancing the response primarily through its synergistic interaction with temperature. The contour lines for q are more symmetrical and uniformly spaced, indicating a smoother gradient and suggesting broader operability within the optimal region.
The contour plot for q identifies a high-capacity adsorption zone (q > 1.1 mg g−1) within a similar operational window, moderate temperatures (around 50 °C), and near-neutral pH values (6.0–6.5). In contrast to the % removal plot, the contour lines here are more concentric and evenly distributed, suggesting that q responds more uniformly to variations in either factor, with less pronounced interaction effects.
The overlap between high-response regions in both contour plots highlights a shared optimal range, demonstrating the feasibility of a unified set of operating conditions that simultaneously maximises % removal and q. These visual representations not only validate the predictive strength of the quadratic regression models but also facilitate the direct identification of robust process windows, making them valuable tools for practical process control and scale-up.
Collectively, these graphical tools reinforce the statistical conclusions derived from the quadratic models and allow for intuitive visualization of factor interactions and response behaviour. The response surfaces confirm the adequacy of model fit and reveal the sensitivity of the process to changes in operating parameters, supporting robust and flexible process design.
Ultimately, the integration of these findings enables effective model-based optimisation and provides mechanistic insight into the adsorption behaviour under variable conditions. The developed models offer a reliable framework for predicting adsorption performance and serve as a foundation for advanced system design and scale-up in dye removal applications.
Although the ANOVA results for the quadratic model indicate that pH was not a statistically significant main factor (p > 0.05), this does not imply that pH lacks influence within the adsorption system. Experimental evidence shows that at pH 6, the ANCSH5 adsorbent achieved higher % removal, which can be attributed to the physicochemical properties of its surface functional groups.
At this pH, the surface of ANCSH5 exposes a greater density of active binding sites, including deprotonated carboxyl groups and aromatic regions capable of π–π stacking with MB+ molecules. These structural features enhance the adsorbent's affinity and capacity for dye uptake, particularly under slightly acidic to neutral conditions.
This mechanistic understanding aligns with the seen experimental enhancement in % removal at pH 6, despite the statistical model assigning low standalone significance to pH. The key lies in the interaction term: ANOVA revealed that the temperature × pH interaction was statistically significant (p < 0.05), indicating that the influence of pH becomes more pronounced when combined with thermal effects.
In contrast, other materials such as ANCSA10 and CPH did not exhibit similar performance at pH 6, likely due to differences in surface functionality and active site availability. This highlights the material-dependent nature of pH sensitivity in adsorption processes.
Thus, the apparent contradiction between statistical significance and experimental trends can be resolved. While pH may not exert a uniform linear influence across the design space, it plays a critical interactive role under specific conditions. The temperature remains the dominant variable, but pH significantly enhances adsorption performance when acting in synergy, particularly in materials like ANCSH5 with favorable surface chemistry.
Importantly, ANCSH5 reached a competitive capacity (100 mg g−1, isotherm-derived) under realistic operating conditions (C0 = 5–30 mg L−1, pH 6, 35 °C, dosage = 4 g L−1), unlike other biomass-derived carbons that report higher qmax values only under extreme concentrations or excessive dosages. This distinctive performance arises from the hemicellulose-rich composition of CPH, which under alkaline activation promotes delignification, partial graphitization, and tailored surface chemistry. Together with validated statistical optimisation, these features establish CPH-derived ANC as a sustainable, low-cost, and industrially competitive adsorbent for dye remediation aligned with circular economy strategies. Overall, the synthesis route not only provides a high-performing adsorbent but also represents a greener alternative by coupling agro-waste valorisation with reduced energy and chemical demands, in contrast to conventional biomass- or fossil-based ACs.
This study did not include high-resolution electron microscopy (HRTEM) with particle-size analysis, density functional theory (DFT) calculations, or standardised ecotoxicity bioassays due to limited access to instrumentation. Nevertheless, the physicochemical characterisation and adsorption performance suggest that the materials may possess nanometric features, as supported by BET, SEM-EDX, XRD, Raman, and FTIR analyses, together with advanced isotherm and kinetic modelling. Additional high-resolution studies will be required to conclusively confirm their nanostructured nature. Future work will integrate HRTEM with quantitative particle sizing, DFT simulations of MB-surface interactions, and OECD-compliant ecotoxicity assays (e.g., Daphnia magna immobilisation and algal growth inhibition) to provide a more comprehensive understanding of structure–property relationships and environmental safety.
| This journal is © The Royal Society of Chemistry 2025 |