Muhammad Irfan Khan*ab,
Suriati Sufian*ab,
Farrukh Hassanc,
Rashid Shamsuddin
d and
Muhammad Farooqe
aDepartment of Chemical Engineering, Universiti Teknologi PETRONAS, Bandar Sri Iskandar, Perak 32610, Malaysia
bCentre of Innovative Nanostructures & Nano Devices (COINN), Institute of Autonomous System, Universiti Teknologi PETRONAS, Seri Iskandar, Perak, Malaysia
cDepartment of Data Science and Artificial intelligence, School of Engineering and Technology, Sunway University, Bandar Sunway, Subang Jaya, Malaysia
dDepartment of Chemical Engineering, Faculty of Engineering, Islamic University of Madinah, 42311, Madinah, Saudi Arabia
eNational Centre of Excellence in Physical Chemistry, University of Peshawar, Peshawar, Pakistan
First published on 22nd January 2025
In this study, a binary composite adsorbent based on activated carbon and phosphoric acid geopolymer foam (ACP) was prepared by combining phosphoric acid geopolymer (PAGP) with activated carbon (AC) and applied for the removal of methylene blue (MB). Activated carbon was thoroughly mixed with a mixture of fly ash and metakaolin in varying ratios, followed by phosphoric acid activation and thermal curing. The ACP adsorbent was characterized using scanning electron microscope (SEM), Fourier transform infrared (FTIR) spectrophotometer, X-ray diffractometer (XRD), surface area analyser (SAP), and thermogravimetric analyser (TGA). Batch analysis was performed to examine the effects of various adsorption parameters including pH (2, 4, 6, 7, 8, and 10), adsorbent dosage (0.06–0.2 g), MB concentration (50–250 mg L−1), contact duration (up to 240 minutes), and temperature (25–55 °C). The ACP with 70% PAGP and 30% AC was found to be the most suitable adsorbent as it maintained its structure and exhibited better MB adsorption. The ACP had a surface area of 47.36 m2 g−1 and a pore size of 5.6 nm and was found to be amorphous in nature. The adsorption equilibrium reached in 240 minutes at pH 7, indicating an efficient adsorption process. The adsorption increased with the initial dye concentration and decreased with the increase in temperature. The ideal parameters for adsorption of MB using ACP include 0.2 g of adsorbent, 25 °C, pH 10, and 240 minutes. The adsorption data fitted well with the Langmuir isotherm, pseudo-second-order (PSO) kinetics model, and three-step intraparticle diffusion (IPD) model. The adsorption capacity calculated using the Langmuir isotherm was 204.8 mg g−1 with an R2 = 0.989. Thermodynamics parameters showed that the adsorption process was exothermic, energetically favourable, and associated with a decrease in entropy. According to the FTIR findings, pH effect, Langmuir isotherm, PSO kinetics, IPD model, and thermodynamics factors, chemisorption is identified as the predominant process. Different machine learning models, i.e., gaussian process regression (GPR), support vector regression (SVR and SVR-rbf), random forest regression (RFR), decision tree regression (DTR) and artificial neural network (ANN), were trained and tested using adsorption capacity and % removal data. The ANN model (random search) demonstrated better performance compared to other models, achieving an R2 value of 0.873 for adsorption capacity and 0.799 for % removal on test data.
Treatment techniques used to screen out MB from the environment include physicochemical (mostly adsorption),8 chemical (using sophisticated oxidation processes),9 and biological (using enzymes and microorganisms) methods.10,11 Adsorption is one of the effective and economic techniques that has the advantage of having realistic working conditions and does not produce secondary contaminants as in the case of chemical technique, e.g. photo catalysis.11,12 The adsorption of dyes is achieved by employing solid adsorbents possessing a large surface area, assembly of active surface functional groups, and robust chemical and physical stability.13 Adsorbents studied for the dyes removal from waste water include carbon-based materials such as activated carbon,14 carbon nanotubes,15 graphene and graphene oxide.16–18 Carbonaceous adsorbents pose a threat to the forestation and plantation, cause thermal pollution, and have poor thermal stability.11 Polymeric adsorbents, a synthetic alternative to carbonaceous adsorbents, are emerging as adsorbents of choice due to their ready availability, commercial production, raw material availability, mechanical stability and tailoring of properties according to the use.19 Despite the advantages offered by polymeric adsorbents, synthetic polymers are increasingly recognized as contemporary pollutants due to the release of toxic chemicals as waste and the production of micro and nanoplastics.20–22 Owing to their very small size, microplastics (MPs) and nanoplastics (NPs) can enter not only to the ecosystem but are detected in human blood and organs, e.g., lungs, liver, gut, and face etc. A litre of bottled water included, on average, almost 2.4 × 105 plastic particles, mostly (90%) in the form of nanoplastics.23 MPs and NPs build up in different organs, causing damage that hinders cellular enzyme action, promotes oxidative stress, or leads to localized inflammation.22,24 Other than microplastics, polymeric products release certain chemicals that are toxic to living organisms, e.g. bisphenol A (BPA) released from plastic container is found to cause early type II diabetes due to insulin resistance.21 Besides diabetes, BPA exposure can cause gene expression leading to the propagation of diseases over generations.25 Owing to the environmental and health risks associated with polymeric products, researchers are attempting to produce environment-friendly and healthy materials to replace their polymeric counterparts.26
Geopolymer and geopolymeric composites have emerged as replacement adsorbents of environmental importance, as they are prepared from waste raw materials, e.g., coal fly ash, steel slag, and ore waste.2 Secondarily, geopolymers are successfully used to adsorb pollutants such as heavy metals,27 uranium,28 dyes,7,29 ammonium and phosphate ions,30 rare earth elements,31 pesticides,32 antibiotics,33 CO234 and surfactants.35 Geopolymers have the potential to partially or completely replace carbonaceous adsorbents for wastewater treatment applications.36 Due to the lower adsorption capacity of geopolymers,37 partial replacement of activated carbon will result in enhanced remediation of pollutants. The binding behaviour of geopolymers helps in maintaining the integrity of the carbon-based adsorbents and offers better thermal stability. In most of the previous studies, alkali-activated geopolymers alone or in combination with secondary materials are used as adsorbents for dye removal.2 In the case of PAGPs, metakaolin is preferably used as a raw material, as fly ash-based PAGPs harden very quickly.38 Due to sharp exotherm and quicker setting, FA-based PAGPS cannot be transformed to foam, and require additional changes in the formulation. The incorporation of metakaolin to regulate setting and exothermic reactions, hydrogen peroxide for foaming, and activated carbon to improve dye adsorption offers a solution to the challenges associated with FA-based PAGPs.
Machine learning (ML), an important component of artificial intelligence (AI), uses algorithms to train models for predictive purposes, eliminating the need for explicit programming for each task.39 The key determinants of AI applications are their simplicity and cost-effectiveness, as they enable exact issue estimate, processing of large amounts of complicated data, and solving highly nonlinear problems that are beyond the reach of empirical equations. Several AI models such as artificial neural network (ANN), multilinear regression (MLR), radial basis function (RBF), support vector machine (SVM), decision tree, random forest regression and artificial neural netweork (ANN) have potentially predicted the adsorption of pollutants from wastewater.39,40 Chong Liu et al. used 12 different machine learning models to predict the adsorption of eight dyes, and the gradient boosting regressor outperformed all models with an R2 value of 0.988 for a total of 627 data points.41 In another study, adsorbents derived from the leaves of Citrus aurantifolia were used for MB adsorption, and the adsorption was predicted using 8 different ML models, among which the linear regression was found to be the best model.42 In most of the previous studies, either % removal or adsorption capacity was used as a dependent variable. As the ML models vary with the type of adsorbents and adsorbates, it is desirable to find an appropriate model for MB adsorption using ACP adsorbents.
In this work, a phosphoric acid based geopolymer foam-activated carbon composite is reported with the aim to partially replace maximum portion of AC by the geopolymer. An activated carbon-phosphoric acid-based geopolymer composites was synthesized in this work and used for MB adsorption. The adsorption parameters were studied using conventional and machine learning models. This work is first of its nature where a phosphoric acid geopolymer foam-based activated carbon composite is prepared and then used for MB removal with the aid of ML tools.
The surface area, pore diameter and pore volume of the adsorbent were determined using a surface area and porosity analyser (ASAP 2020, Micromeritics) by N2 adsorption and desorption methods. The samples were degassed at 100 °C for 240 minutes. A simultaneous thermal analyzer (STA 6000, PerkinElmer) was used for the thermogravimetric analysis of the ACP sample. About 5 mg sample was thermally analysed under an inert (N2) atmosphere, in the temperature range of 30 °C to 700 °C at a heating rate of 10 °C min−1.
![]() | (1) |
![]() | (2) |
pH | Dosage (g) | Conc. (mg L−1) | Time (min) | Temp (°C) | |
---|---|---|---|---|---|
Effect of pH | 2, 4, 6, 7, 8, 10 | 0.2 | 250 | 240 | 25 |
Effect of dosage | 7 | 0.06, 0.08, 0.10, 0.12, 0.15, 0.20 | 250 | 240 | 25 |
Effect of dye conc. | 7 | 0.2 | 50, 100, 150, 200, 250 | 240 | 25 |
Effect of time | 7 | 0.2 | 100, 150, 200 | 10, 20, 30, 40, 60, 120, 180, 240 | 25 |
Effect of temp. | 7 | 0.2 | 200 | 240 | 25, 35, 45, 55 |
![]() | (3) |
![]() | (4) |
Qe = B![]() ![]() ![]() ![]() | (5) |
In eqn (3) (Langmuir isotherm), the maximum adsorption capacity and the adsorption energy are denoted by Qm, and KL, respectively. In the Freundlich equation, the adsorption intensity is indicated by KF, whereas n is the Freundlich constant. Furthermore, B is the Temkin constant for the sorption heat (J mol−1) and A is the Temkin isotherm constant (L g−1). In the isotherm investigation, the best fit (highest R2) was used as the benchmark to determine which model best describes the adsorption process. RL is an additional parameter that is derived from the Langmuir isotherm and is computed using eqn (6):
![]() | (6) |
The favourability of adsorption is predicted by the RL value based on the Langmuir isotherm: when RL = 0, all adsorption sites are occupied and the adsorption is irreversible. For favourable adsorption, 0 < RL < 1, and unfavourable adsorption is represented by RL > 1.45
Kinetic parameters were determined by analysing the data derived from the adsorption of MB against time. Using Lagergren pseudo-first-order (PFO) and pseudo-second-order (PSO) eqn (7) and (8), the kinetics of the MB adsorption process was evaluated.46 Additionally, intraparticle diffusion (IPD) and liquid film diffusion (LFD) models' equations (eqn (9) and (10)) were applied to predict the adsorption mechanism and the diffusion mode.46,47 In these equations, the adsorption capacity at a certain time t is represented by Qt, t is the time interval, the intercept of the IPD model is denoted by C, and F is equal to Qt/Qe. R2 was used to describe the best-fitted kinetics model:
![]() | (7) |
![]() | (8) |
![]() | (9) |
ln(1 − F) = kFT | (10) |
Lastly, the thermodynamic parameters, which include the enthalpy change (ΔH), change in free energy (ΔG), and change in entropy (ΔS), were determined using eqn (11)–(13):48
![]() | (11) |
![]() | (12) |
ΔG = ΔH − TΔS | (13) |
Independent variable | Dependent variable/response | |||
---|---|---|---|---|
MB concentration | Ads. time | pH | % removal | Ads. capacity |
150 | 10 | 4 | 85.32 | 31.99 |
150 | 20 | 4 | 95.60 | 35.85 |
150 | 30 | 4 | 97.75 | 36.66 |
150 | 40 | 4 | 98.44 | 36.92 |
150 | 60 | 4 | 98.61 | 36.98 |
150 | 90 | 4 | 98.67 | 37.00 |
150 | 120 | 4 | 99.47 | 37.30 |
150 | 180 | 4 | 99.84 | 37.44 |
150 | 240 | 4 | 99.99 | 37.50 |
150 | 10 | 7 | 98.35 | 36.88 |
150 | 20 | 7 | 99.02 | 37.13 |
150 | 30 | 7 | 99.37 | 37.26 |
150 | 40 | 7 | 99.38 | 37.27 |
150 | 60 | 7 | 99.41 | 37.28 |
150 | 90 | 7 | 99.54 | 37.33 |
150 | 120 | 7 | 99.60 | 37.35 |
150 | 180 | 7 | 99.84 | 37.44 |
150 | 240 | 7 | 99.90 | 37.46 |
150 | 10 | 10 | 93.73 | 35.15 |
150 | 20 | 10 | 95.96 | 35.98 |
150 | 30 | 10 | 98.04 | 36.77 |
150 | 40 | 10 | 98.62 | 36.98 |
150 | 60 | 10 | 99.22 | 37.21 |
150 | 90 | 10 | 99.04 | 37.14 |
150 | 120 | 10 | 99.74 | 37.40 |
150 | 180 | 10 | 99.94 | 37.48 |
150 | 240 | 10 | 99.99 | 37.49 |
200 | 10 | 4 | 77.50 | 38.75 |
200 | 20 | 4 | 92.03 | 46.01 |
200 | 30 | 4 | 95.60 | 47.80 |
200 | 40 | 4 | 97.21 | 48.61 |
200 | 60 | 4 | 98.62 | 49.31 |
200 | 90 | 4 | 98.90 | 49.45 |
200 | 120 | 4 | 99.64 | 49.82 |
200 | 180 | 4 | 99.69 | 49.85 |
200 | 240 | 4 | 99.92 | 49.96 |
200 | 10 | 7 | 92.23 | 46.11 |
200 | 20 | 7 | 96.29 | 48.15 |
200 | 30 | 7 | 98.09 | 49.04 |
200 | 40 | 7 | 98.28 | 49.14 |
200 | 60 | 7 | 99.24 | 49.62 |
200 | 90 | 7 | 99.51 | 49.75 |
200 | 120 | 7 | 99.51 | 49.76 |
200 | 180 | 7 | 99.63 | 49.82 |
200 | 240 | 7 | 99.69 | 49.85 |
200 | 10 | 10 | 84.24 | 42.12 |
200 | 20 | 10 | 96.80 | 48.40 |
200 | 30 | 10 | 97.73 | 48.87 |
200 | 40 | 10 | 98.25 | 49.12 |
200 | 60 | 10 | 98.77 | 49.38 |
200 | 90 | 10 | 99.02 | 49.51 |
200 | 120 | 10 | 99.72 | 49.86 |
200 | 180 | 10 | 99.81 | 49.91 |
200 | 240 | 10 | 99.96 | 49.98 |
250 | 10 | 4 | 76.14 | 47.59 |
250 | 20 | 4 | 84.54 | 52.84 |
250 | 30 | 4 | 88.53 | 55.33 |
250 | 40 | 4 | 89.36 | 55.85 |
250 | 60 | 4 | 94.38 | 58.99 |
250 | 90 | 4 | 96.76 | 60.47 |
250 | 120 | 4 | 98.01 | 61.25 |
250 | 180 | 4 | 98.42 | 61.51 |
250 | 240 | 4 | 99.17 | 61.98 |
250 | 10 | 7 | 70.86 | 44.29 |
250 | 20 | 7 | 77.65 | 48.53 |
250 | 30 | 7 | 91.85 | 57.41 |
250 | 40 | 7 | 93.75 | 58.60 |
250 | 60 | 7 | 96.61 | 60.38 |
250 | 90 | 7 | 98.17 | 61.36 |
250 | 120 | 7 | 98.95 | 61.84 |
250 | 180 | 7 | 99.11 | 61.95 |
250 | 240 | 7 | 99.64 | 62.27 |
250 | 10 | 10 | 66.54 | 41.59 |
250 | 20 | 10 | 81.68 | 51.05 |
250 | 30 | 10 | 93.77 | 58.61 |
250 | 40 | 10 | 93.99 | 58.75 |
250 | 60 | 10 | 97.39 | 60.87 |
250 | 90 | 10 | 98.52 | 61.58 |
250 | 120 | 10 | 98.97 | 61.86 |
250 | 180 | 10 | 99.38 | 62.12 |
250 | 240 | 10 | 99.69 | 62.31 |
![]() | ||
Fig. 2 (a) FTIR analysis of the PAGP and ACP adsorbents. Deconvolution of the geopolymer peak in (b) PAGP and (c) ACP. |
A slight shift of the geopolymer peak to a higher wavenumber in the ACP sample compared to PAGP indicates interaction between PAGP and AC. An upsurge shift in the frequency band at 1045 cm−1 for the PAGP sample to 1049 cm−1 for the ACP and a subsequent decrease in peak intensity and peak area indicates that the addition of activated carbon hindered geopolymer formation. This phenomenon may be caused by the presence of attractive forces between PAGP and AC, e.g., hydrogen bonding, surface group interactions and impact of molecular environment. Moreover, part of the phosphoric acid may be adsorbed by AC reducing its activation behaviour.
The powder X-ray diffraction analysis of the PAGP and ACP, in the 2 theta range of 2–80°, is presented in Fig. 3. Analysis of the XRD data reveals that both samples are mostly amorphous with traces of some crystalline phases. A broad hump in the 20–27° 2 theta range is particularly noticeable in Fig. 4 and is indicative of the material's amorphous nature.38 In addition to the amorphous material, the quartzite phase is very prominent as evidenced by the peaks at 20.8°, 26.5°, 50.2°, 60.1°, and 68.3°.29 Both ACP and PAGP show the existence of brushite and mullite phases. The amorphous material and brushite phase existence as indicated by the PXRD patterns represents the formation of PAGP in both samples, whereas quartzite and mullite phases manifest the unreacted FA. The XRD is in close agreement with the FTIR analysis, as it also indicates the presence of unreacted FA.
![]() | ||
Fig. 4 (a–c) SEM micrographs of ACP of different points at different magnifications. (d) EDX analysis of ACP. |
The SEM micrographs of the ACP at different resolutions (3000×–5000×) are presented in Fig. 4(a–c). Two types of morphologies are visualized in the micrographs, i.e., geopolymer and fly ash particles embedded in the geopolymer matrix with the sizes in the range of 5 μm to 15 μm. Small plate-like morphologies observed in Fig. 4c are representing activated carbon. The EDX analysis is shown in Fig. 4d, which indicates that the main constituents of ACP are Al, Ca, Si, S, P, C, and O.
The formation of geopolymers, homogeneity of the geopolymer and AC, and the interfacial transition zone between the geopolymer and AC can be established from the SEM analysis.29 The homogenous geopolymer matrix indicated the successful formation of PAGP in the presence of activated carbon. Large-sized unreacted and partially reacted fly ash particles demonstrated that large particles are mostly crystalline and less reactive.51 Mechano-chemical activation using ball milling or pre-screening improves the geopolymer content and reduces the unreacted fly ash quantity.52 The SEM images agree closely with the XRD and FTIR analysis discussed earlier. The EDX analysis presents that a P to Al ratio of 1 is maintained by ACP. The high % of C is linked with the AC in the sample and the carbon tape used for mounting sample to aluminum stub.
The nitrogen adsorption–desorption isotherm obtained by the BET surface area analysis is depicted in Fig. 5. ACP exhibited a surface area of 47.36 m2 g−1, an adsorption pore diameter of 5.39 nm, and a cumulative pore volume of 0.0206 cm3 g−1 (ascertained by BJH adsorption within the pore size range of 1.700 to 300 nm). A hysteresis loop, which is indicative of a type IV isotherm in the IUPAC classification, is shown by the isotherm.53 The findings showed that the combination of the geopolymer and AC resulted in a mesoporous material having a considerably higher surface area than that of geopolymers alone.29 A pore diameter of 5.39 nm also supports the mesoporous nature of the ACP being in the range of 2–50 nm.
The TGA study findings for the ACP are presented in Fig. 6. The analysis was performed in a N2 atmosphere at a heating rate of 10 °C min−1, encompassing in the temperature range of 25 to 700 °C. Across the three apparent steps, there is a 16% mass loss in the first step, i.e., below 100 °C followed by the 2nd mass loss in the temperature range of 100–500 °C, and a steady mass loss of 4.4% is observed, whereas 4% mass loss is recorded above 500 °C. A cumulative mass loss of 24.4% is attained during TG analysis.
The first step is a representation of the evaporation of the physically adsorbed water molecules (weakly linked with the solid surface) on the surface of the ACP. The less steep step 2 in the TGA curve from 200 to 600 °C indicates the breakdown of the germinal and vicinal silanol groups, whereas mass loss from 600 to 700 °C is the outcome of the isolated silanol groups' breakdown.54 TG analysis reveals that ACP exhibits good thermal stability up to 700 °C.
The effect of pH on the adsorption of MB, in terms of % removal and adsorption capacity, on the ACP adsorbent is presented in Fig. 8. The % removal achieved ranged between 97.98 and 99.50% at different pH values with the highest removal achieved at pH 10. There is a slight difference among the %R at different pH values, which represent that ACP is effective at all pH values. It is necessary to mention that the final pH of all the samples, after 4 hours of shaking, was recorded in the range of pH 6 to pH 8, representing that ACP has normalized the solution pH to neutral pH.
Similar to % removal, the adsorption capacity has shown a slight variation as a function of pH and remained in the range of 61.2–62.1 mg g−1, establishing that the adsorbent is equally effective at acidic, neutral, and alkaline pH values. The maximum adsorption capacity of 62.11 mg g−1 was achieved at pH 10. The uniform adsorption capacity and % removal at different pH levels are linked to the chemistry of PAGPs, which tends to adsorb maximum amount of MB at pH 10.
Fig. 9 presents the percent removal and adsorption capacity of MB as a function of adsorbent dosage at a fixed MB concentration (250 mg L−1) and volume (50 mL). The results reveal that 15.99, 28.77, 37.53, 57.15, 72.27 and 99.69% of MB was adsorbed at the adsorbent doses of 0.06, 0.08, 0.10, 0.12, 0.15 and 0.20 g/50 ml of MB solution, respectively. The adsorption capacity obtained for 0.06 to 0.20 g dosage ranged from 33.3 to 62.3 mg g−1. Overall, an upsurge in the % removal and adsorption capacity is attained with the increase in adsorbent dosage, and the trend is explained on the grounds that with more adsorbent, there is an increase in adsorption sites due to the presence of more adsorbent materials. The adsorption capacity increases with the higher adsorbent doses due to increased surface area, active sites availability, and interaction possibilities. This enhances the efficiency and leads to multilayer adsorption.15
Fig. 10 illustrates the graphical presentation, depicting the influence of MB concentration on both the %R and adsorption capacity of the ACP adsorbent. The removal efficiency of ACP exhibits values of 99.88%, 99.87%, 99.88%, 99.82%, and 99.78% for MB concentrations of 50 mg L−1, 100 mg L−1, 150 mg L−1, 200 mg L−1, and 250 mg L−1, respectively.
However, the recorded adsorption capacity manifests values of 12.48, 24.97, 37.45, 49.91, and 62.37 mg g−1 for the MB concentrations of 50, 100, 150, 200 and 250 mg L−1, respectively. Remarkably, a negligible decreasing trend in terms of %R and a prominent increasing trend in adsorption capacity are evident in this experimental investigation. These findings feature a concentration-dependent behaviour of the MB adsorption process and contribute valuable insights for optimizing the performance of ACP in MB removal applications.
A concentration-dependent adsorption phenomenon is suggested by the rise in methylene blue's adsorption capacity using ACP that is observed as the dye concentration increased from 50 to 250 mg L−1. The outcome suggests that more MB molecules are successfully adsorbed onto the ACP surface as the initial dye concentration increases. This is probably because there is a stronger driving force for mass transfer. The slight decrease in %R may be explained by the concentration effects, i.e., at high concentrations, the MB molecules have affinity between them.
Studying the adsorption process's kinetics and thermodynamics is necessary, in which time and temperature play a role. MB adsorption on ACP adsorbent is displayed as a function of time for 240 minutes, as shown in Fig. 11(a and b). In the plots depicting % removal vs. time (%R vs. t) and adsorption capacity vs. time (Qt vs. t), equilibrium is achieved within 30 minutes for solutions with concentrations of 150 mg L−1 and 200 mg L−1. In contrast, 250 mg L−1 solution requires 60 minutes to reach equilibrium. Irrespective of the different equilibrium times, 70–90% of MB has been adsorbed in first 10 minutes. This quicker adsorption compared to geopolymers and AC-based adsorbents is the unique feature of the ACP, as most of the geopolymeric adsorbents take at least 60–240 minutes to reach the equilibrium point.8,55 This quicker adsorption of MB is achieved due to the charged structure of PAGP, high surface area of ACP, foaming property by H2O2 and additional activation of AC by H3PO4 used for PAGP formation. Phosphoric acid treatment chemically activates AC by introducing additional pores and enhancing surface area, enhancing adsorption capacity and reactivity. This modification also influences the functional groups on the surface, contributing to improved MB capture and removal, particularly in applications of water purification.
![]() | ||
Fig. 11 Time influence on MB adsorption by ACP at initial concentrations (Ci) of 150, 200 and 250 mg L−1: (a) % removal (b) adsorption capacity. |
The influence of temperature on the adsorptive removal of MB using ACP adsorbent is illustrated in Fig. 12. The figure distinctly demonstrates a decreasing trend in MB adsorption as the temperature rises from 25 °C to 35 °C, 45 °C and 55 °C. The adsorption percentages were 99.68%, 99.22%, 97.83% and 96.65% at 25 °C, 35 °C, 45 °C and 55 °C, respectively. Mirroring the trend observed in % removal, the adsorption capacity of ACP exhibited a declining pattern (49.8, 49.6, 48.9, and 48.3 mg g−1) with the increase in temperature. These results indicate the involvement of exothermic adsorption, and the decrease in adsorption with elevated temperature suggests the extent of this exothermic adsorption process.48
Additionally, increased temperatures disturb the equilibrium between adsorption and desorption by supplying increased thermal energy, causing an accelerated desorption of methylene blue from the adsorbent's surface, hence reducing adsorption. Furthermore, elevated temperatures may induce molecular alterations in both the adsorbent and methylene blue, further impacting their interaction.56 A similar trend has been reported in geopolymer-based adsorbents for MB removal.2,29,57
![]() | ||
Fig. 13 Adsorption isotherms of MB using ACP: (a) Langmuir isotherm, (b) Freundlich isotherm, and (c) Temkin isotherm. (d) Variation in RL with concentration. |
The degree of fitting (R2) of the linear fit of the data from experiments was used to determine the optimal isotherm from the equilibrium concentration (Ce) vs. adsorption capacity (Qe) data at different concentrations of MB. As shown in Table 3, the degree of fitting of Langmuir isotherm is higher than those of the Freundlich and Temkin isotherms. The linear fits of the Freundlich and Temkin isotherms yielded lower degrees of fitting, i.e., 0.947 and 0.984 compared to the Langmuir isotherm (0.989). Qm, KL, and R2 values of 204.08 mg g−1, 1.02 L mg−1, and 0.989, respectively, were obtained by the Langmuir isotherm.
Isotherm | Parameters | Values |
---|---|---|
Langmuir | Qm (mg g−1) | 204.08 |
KL (L mg−1) | 1.02 | |
RL | 0.019 | |
R2 | 0.989 | |
Freundlich | N | 1.2832 |
KF | 112.415 | |
R2 | 0.947 | |
Temkin | A (L g−1) | 24.3772 |
B (J mol−1) | 23.68 | |
R2 | 0.984 |
In the Langmuir isotherm, it is believed that adsorption occurs on the uniform surfaces of the adsorbent material, leading to a monolayer adsorption process. Furthermore, the reversibility of the adsorption and desorption processes is postulated and the adsorption sites on adsorbent are homogeneous, meaning that their energies are all identical.29,59,60 Additionally, the model suggests a maximum adsorption capacity, above which no more adsorption occurs. The adsorption is a reversible process and ceases once a site is occupied, preventing any contact between the molecules of the adsorbate. The favourability of the adsorption is further supported by the RL (Fig. 13d) value calculated using eqn (6) as it lies between 0 and 1. The extent of fitting of the Freundlich and Temkin isotherms (R2 = 0.947 and 0.984) is reasonably closer to the Langmuir isotherm, representing that along with homogenous adsorption, heterogeneous adsorption occurs.
The value of n obtained using the Freundlich isotherm is 1.283, surpassing the threshold of 1. A linear adsorption process with uniform adsorption is indicated by a value of n = 1. Conversely, if n is greater than 1, it means that the adsorbent's adsorption capacity increases more sharply as the concentration of MB in the solution increases. This behaviour is frequently linked to the strong interactions between the adsorbent and the adsorbate as well as the occurrence of multilayer adsorption. It provides more evidence for the existence of multilayer adsorption and chemisorption processes. Furthermore, unfavourable and irreversible adsorption processes are represented by n < 1 and n = 0.61
Qe,exp (mg g−1) | Qe,Cal (mg g−1) | k1(PFO) (min−1) | k2(PSO) (min−1) | kIPD (mg g−1 s−1/2) | kLFD (mg g−1 min−1) | R2 | |
---|---|---|---|---|---|---|---|
PFO (150 mg L−1) | 37.46 | 4.06 | 0.0001 | 0.630 | |||
PFO (200 mg L−1) | 49.84 | 3.15 | 0.0007 | 0.419 | |||
PFO (250 mg L−1) | 62.27 | 3.08 | 0.0248 | 0.939 | |||
PSO (150 mg L−1) | 37.46 | 37.45 | 0.106 | 0.999 | |||
PSO (200 mg L−1) | 49.84 | 50.00 | 0.030 | 0.999 | |||
PSO (250 mg L−1) | 62.27 | 63.29 | 0.004 | 0.999 | |||
IPD step 01 | 49.84 | 0.1652 | 0.979 | ||||
IPD step 02 | 49.84 | 0.0245 | 0.869 | ||||
IPD step 3 | 49.84 | 0.0025 | 0.981 | ||||
LFD (150 mg L−1) | 37.46 | 0.00007 | 0.630 | ||||
LFD (200 mg L−1) | 49.84 | 0.0007 | 0.419 | ||||
LFD (250 mg L−1) | 62.27 | 0.0250 | 0.939 |
As indicated in Table 4, the PFO model exhibited R2 values of 0.630, 0.419, and 0.939 for MB solutions having concentrations of 150, 200, and 250 mg L−1, respectively. In contrast, the PSO model demonstrated a robust fit with an R2 value of 0.999 across all concentrations. Beyond the R2 values, the adsorption capacity calculated using the PFO model falls within the range of 3.06–4.08 mg g−1, which is nearly ten times lower than the experimentally determined adsorption capacity. On the contrary, noteworthy agreement between the estimated and actual adsorption capacities is evident in the kinetics data of the PSO model strengthening the conclusion that the kinetics process followed by ACP is the PSO model at each concentration. KPSO, rate of adsorption of MB by ACP, showed a decreasing trend with the increase in MB concentration, representing that the rate of adsorption is decreasing with the surge in MB concentration.
Several important aspects of the adsorption process are implied by the adherence of adsorption kinetics to the PSO model. It is proposed that chemisorption is more prevalent than physical adsorption, meaning that chemical bonds are formed during the contact between the MB and the ACP, resulting in a more strong and focused binding. Emphasizing the uniformity of the adsorption sites and their attraction for the MB, the model also assumes a homogenous surface with uniform adsorption sites. On the surface, MB molecules are organized in a monolayer according to the preference for a single-layer adsorption process. Additionally, according to the PSO model, the chemical interaction determines the rate of adsorption, making it the process's rate-limiting phase.29,62
The drop in k value seen during the PSO kinetics of MB by increasing the concentration can be ascribed by the quick saturation of adsorption sites at higher concentrations and increased competition amongst MB molecules for these sites. The surface saturation restricts the number of adsorption sites available, which lowers the rate constant. A lower k is the outcome of increased competition amongst MB molecules, which reduces the overall efficiency of the adsorption process. Furthermore, there may be possible changes in the way that MB and the adsorbent surface interact, diffusion restrictions, and site heterogeneity.
Fig. 14c and d show the IPD and the LFD models applied to adsorption kinetics data, respectively. The IPD model applied to the whole dataset produced a linear fit with an R2 value of 0.556, suggesting that the single-step IPD model was not the main rate-limiting mechanism. Similarly, the LFD model did not fit well to the experimental data producing a lower R2 value.47,63,64 Furthermore, a three-step IPD model produced R2 values of 0.979, 0.869 and 0.981 for step 1, 2 and 3, respectively. The rate constant (k) exhibited a decrement from step 1 to step 2 and a further increment in step 3, indicating a swifter adsorption in the initial step that varies over the time. This decrease signifies that most of the adsorption occurs within the first 30 minutes, highlighting a reduction in adsorption efficiency as the process time advances. Moreover, the non-zero intercept value showed that intraparticle diffusion was not the sole phase in the adsorption process.65 These findings showed that intraparticle diffusion was not the sole phase in the adsorption process that limited the speed, and the model's linear curve did not pass through the origin (0,0). The findings lead to the conclusion that neither the single-step IPD model nor the LFD model constitutes the primary mechanism. Instead, the three-step IPD model provides a superior fit to the experimental data.
In the prior research related to the adsorption of MB onto PAGPs and alkali-activated geopolymers (AAGPs), a dominant trend emerges with the consistent adoption of the PSO model.43 This signifies the predominant role played by the ionic structure inherent in geopolymers, whether in the form of PAGPs or AAGPs, in driving the adsorption process through predominant ionic interactions. The widespread utilization of the PSO model across diverse studies underlines a consideration of the fundamental mechanisms governing MB adsorption onto geopolymers, shedding light on the pivotal influence of the ionic nature of geopolymers in facilitating adsorption interactions.
![]() | ||
Fig. 15 Plot of the linear correlation between the van ‘t Hoff equation and temperature used for the adsorption of MB dye on ACP. |
T (°C) | ΔH kJ mol−1 | ΔG kJ mol−1 | ΔS J mol−1 K−1 | R2 |
---|---|---|---|---|
25 | −67.443 | −10.742 | −190.174 | 0.983 |
35 | −8.840 | |||
45 | −6.939 | |||
55 | −5.037 |
The thermodynamic parameters that are computed for the adsorption of MB on ACP provide information about the characteristics of the adsorption process. The enthalpy change is negative (ΔH = −67.443 kJ mol−1), suggesting that heat is released during the adsorption process and the adsorption process is exothermic. Furthermore, the adsorption process on ACP appears to be spontaneous based on the negative change in Gibbs free energy (ΔG = −10.742 to −5.037 kJ mol−1), implying that there is an energy favourability to the procedure. As the MB adsorption on ACP is more favourable at a lower ΔG value, therefore lower temperature is better for MB adsorption.65 A reduction in randomness at the interface is shown by the negative entropy change (ΔS = −190.174 J mol−1 K−1), which represents a more ordered adsorption system after adsorption. All parameters considered, these thermodynamic parameters offer thorough insights into the spontaneity and energetics of the MB-ACP adsorption, which are essential for comprehending and refining the adsorption process.48,62,64
![]() | ||
Fig. 16 (a) Re-use of ACP for 5 times (MB conc. = 200 mg L−1, V = 50 ml and dosage = 0.2 g) (b) % loss in adsorption with cycles. |
Combining characterization, kinetics, isotherms and thermodynamics, the MB adsorption on ACP is a combination of physisorption and chemisorption owing to multiple types of functional groups of PAGP and AC.
Finally, following three types of interactions could be responsible for MB adsorption.
(1) PAGP has negatively charged aluminium and positively charged phosphorus atoms and these atoms have strong interactions with the cationic MB molecules, as shown in reaction Scheme 1.29,71 As both PAGP and MB are having charged centres, this interaction will be the dominant mechanism.
(2) The second type of interaction includes the stronger attractions between MB and various functional groups on the surface of the activated carbon, e.g. the carboxylate, hydroxyl and carbonyl groups, as given in reaction Scheme 2. Activated carbon has hydroxyl, carboxyl, carbonyl, vinyl, ethereal and anhydride types of functional groups that attract MB molecules using hydrogen bonding, dipole–dipole interactions, and van der Waals forces etc.10
(3) The third type of interaction includes weak non bonded intermolecular forces between MB and adsorbents, e.g. London forces and dipole-induced dipole forces.10
Adsorbent | Activator | Isotherm | Kinetics | Energetics | Mechanism | Q (mg g−1) | Ref. |
---|---|---|---|---|---|---|---|
FA-GP | Alkali | Freundlich | PSO | Not discussed | Chemisorption | 79.81 | 8 |
MK-GP | Alkali | Langmuir | PSO | Not discussed | Chemisorption | 39.5 | 60 |
MK-GP | H3PO4 | Langmuir | PSO | Exothermic | Chemisorption | 3.01 | 29 |
MK-GP foam | Alkali | Freundlich | PSO | Endothermic | Chemisorption | 12.5 | 7 |
MK-GP foam | Alkali | Langmuir | PSO | Endothermic | Chemisorption | 40.0 | 72 |
Graphitic carbon nitride/GP | Alkali | Langmuir | PSO | Exothermic | Chemisorption | 170.9 | 73 |
FAGP/TiO2 | Alkali | Langmuir | PSO | Exothermic | Chemisorption | 103.19 | 37 |
ACP | H3PO4 | Langmuir | PSO | Exothermic | Chemisorption | 204.08 | This study |
Model type | RMSE (Train) | R2 (Train) | MAE (Train) | RMSE (Test) | R2 (Test) | MAE (Test) |
---|---|---|---|---|---|---|
Gaussian process regression (rotational quadratic) | 2.544 | 0.939 | 1.871 | 2.889 | 0.858 | 1.931 |
Support vector regression (kernel = ‘poly’) | 1.889 | 0.967 | 1.266 | 5.325 | 0.518 | 3.343 |
Support vector regression (kernel = ‘rbf’) | 7.661 | 0.453 | 6.702 | 6.099 | 0.369 | 4.914 |
Random forest regression (n_estimators = 10) | 0.577 | 0.997 | 0.338 | 3.499 | 0.792 | 2.001 |
Decision tree regression (max_depth = 10) | 0.8089 | 0.994 | 0.456 | 2.824 | 0.865 | 1.624 |
Artificial neural network (ANN) | 1.323 | 0.984 | 1.107 | 2.737 | 0.873 | 1.912 |
Model type | RMSE (Train) | R2 (Train) | MAE (Train) | RMSE (Test) | R2 (Test) | MAE (Test) |
---|---|---|---|---|---|---|
Gaussian process regression (rotational quadratic) | 2.489 | 0.679 | 2.0316 | 7.932 | 0.326 | 4.660 |
Support vector regression (kernel = ‘poly’) | 2.639 | 0.638 | 1.344 | 8.609 | 0.205 | 4.443 |
Support vector regression (kernel = ‘rbf’) | 3.539 | 0.348 | 1.521 | 9.815 | 0.0325 | 5.094 |
Random forest regression (n_estimators = 10) | 0.660 | 0.977 | 0.448 | 6.040 | 0.608 | 3.067 |
Decision tree regression (max_depth = 10) | 2.275 | 0.730 | 1.137 | 5.596 | 0.664 | 3.194 |
Artificial neural network (ANN) | 1.778 | 0.835 | 1.394 | 4.320 | 0.799 | 3.29 |
The choice of kernel function significantly influenced the performance of support vector regression (SVR) models. For example, when employed a support vector regression (SVR) model with a polynomial kernel, an R2 value of 0.967 was attained during training for adsorption capacity data. However, when using rbf kernel, the R2 value decreased to 0.453. Similarly, for % removal prediction, the R2 value decreased from 0.638 to 0.348.
To ensure the optimal performance and minimize errors, hyperparameter optimization was conducted for the five algorithms applied to this data.
Optimizer | Interactions/grids | Error index | |||||
---|---|---|---|---|---|---|---|
RMSE | R2 | MAE | |||||
Train | Test | Train | Test | Train | Test | ||
Gaussian process regression (GPR) | |||||||
Bayesian | 30 | 3.324 | 2.98 | 0.858 | 0.858 | 2.37 | 2.19 |
40 | 2.685 | 2.889 | 0.858 | 0.858 | 1.931 | 1.931 | |
50 | 2.685 | 2.889 | 0.858 | 0.858 | 1.9311 | 1.931 | |
Grid search | 2 | 2.544 | 2.889 | 0.939 | 0.858 | 1.871 | 1.931 |
3 | 2.544 | 2.889 | 0.939 | 0.858 | 1.871 | 1.931 | |
5 | 2.544 | 2.889 | 0.939 | 0.858 | 1.871 | 1.931 | |
Random search | 30 | 2.646 | 2.932 | 0.858 | 0.858 | 1.903 | 2.030 |
40 | 2.658 | 2.935 | 0.858 | 0.858 | 1.903 | 1.903 | |
![]() |
|||||||
Random forest regressor (RFR) | |||||||
Bayesian | 30 | 0.561 | 6.432 | 0.973 | 0.532 | 0.674 | 3.984 |
40 | 0.561 | 6.432 | 0.973 | 0.532 | 0.456 | 3.984 | |
50 | 0.557 | 3.449 | 0.997 | 0.792 | 0.338 | 2.001 | |
Grid search | 2 | 0.572 | 5.640 | 0.982 | 0.674 | 0.456 | 2.348 |
3 | 0.572 | 6.432 | 0.973 | 0.674 | 0.456 | 2.348 | |
5 | 0.572 | 6.432 | 0.973 | 0.674 | 0.674 | 3.987 | |
Random search | 30 | 0.561 | 6.432 | 0.973 | 0.532 | 0.674 | 3.987 |
40 | 0.561 | 6.432 | 0.973 | 0.532 | 0.674 | 2.348 | |
![]() |
|||||||
Decision tree | |||||||
Bayesian | 30 | 0.823 | 5.632 | 0.981 | 0.664 | 0.562 | 3.712 |
40 | 0.823 | 5.632 | 0.981 | 0.664 | 0.562 | 3.712 | |
50 | 0.823 | 5.632 | 0.981 | 0.664 | 0.562 | 3.634 | |
Grid search | 2 | 0.823 | 5.632 | 0.981 | 0.664 | 0.562 | 3.634 |
3 | 0.809 | 2.824 | 0.994 | 0.865 | 0.456 | 1.624 | |
5 | 0.836 | 4.640 | 0.981 | 0.817 | 0.569 | 2.520 | |
Random search | 30 | 0.823 | 5.632 | 0.981 | 0.664 | 0.562 | 3.624 |
40 | 0.823 | 5.632 | 0.981 | 0.664 | 0.562 | 3.624 | |
![]() |
|||||||
Artificial neural network (ANN) | |||||||
Bayesian | 30 | 2.734 | 2.897 | 0.978 | 0.817 | 3.532 | 2.154 |
40 | 1.997 | 4.961 | 0.978 | 0.864 | 2.928 | 4.234 | |
50 | 11.323 | 9.024 | 0.589 | 0.563 | 8.678 | 7.523 | |
Grid search | 2 | 1.321 | 3.921 | 0.984 | 0.824 | 3.425 | 3.456 |
3 | 1.422 | 3.921 | 0.984 | 0.824 | 2.228 | 2.228 | |
5 | 1.422 | 3.921 | 0.984 | 0.824 | 2.228 | 2.228 | |
Random search | 30 | 1.323 | 2.737 | 0.984 | 0.873 | 1.107 | 1.912 |
40 | 1.712 | 2.862 | 0.984 | 0.814 | 2.234 | 2.453 | |
![]() |
|||||||
Support vector machine (SVM) | |||||||
Bayesian | 30 | 1.889 | 5.325 | 0.967 | 0.518 | 1.266 | 3.343 |
Optimizer | Interactions/grids | Error index | |||||
---|---|---|---|---|---|---|---|
RMSE | R2 | MAE | |||||
Train | Test | Train | Test | Train | Test | ||
Gaussian process regression (GPR) | |||||||
Bayesian | 30 | 4.442 | 9.014 | 0.523 | 0.321 | 2.042 | 5.453 |
40 | 4.442 | 8.005 | 0.632 | 0.321 | 2.053 | 5.453 | |
50 | 3.564 | 7.664 | 0.632 | 0.321 | 2.053 | 4.675 | |
Grid search | 2 | 3.564 | 7.764 | 0.632 | 0.326 | 2.053 | 4.675 |
3 | 2.489 | 7.932 | 0.679 | 0.326 | 2.031 | 4.660 | |
5 | 2.489 | 7.993 | 0.679 | 0.326 | 2.031 | 4.660 | |
Random search | 30 | 2.489 | 7.993 | 0.545 | 0.326 | 2.134 | 4.660 |
40 | 2.642 | 7.993 | 0.545 | 0.326 | 2.134 | 4.660 | |
![]() |
|||||||
Random forest regression (RFR) | |||||||
Bayesian | 30 | 0.775 | 7.222 | 0.962 | 0.552 | 0.534 | 4.234 |
40 | 0.775 | 7.222 | 0.972 | 0.567 | 0.467 | 3.034 | |
50 | 0.775 | 6.132 | 0.972 | 0.559 | 0.467 | 3.034 | |
Grid search | 2 | 0.779 | 6.132 | 0.983 | 0.554 | 0.448 | 3.034 |
3 | 0.771 | 6.132 | 0.983 | 0.554 | 0.448 | 4.165 | |
5 | 0.771 | 6.132 | 0.983 | 0.559 | 0.448 | 3.067 | |
Random search | 30 | 0.660 | 6.040 | 0.997 | 0.608 | 0.448 | 3.067 |
40 | 0.660 | 6.040 | 0.992 | 0.603 | 0.459 | 3.192 | |
![]() |
|||||||
Decision tree | |||||||
Bayesian | 30 | 2.281 | 5.596 | 0.721 | 0.663 | 1.229 | 3.200 |
40 | 2.275 | 5.596 | 0.721 | 0.667 | 1.137 | 3.204 | |
50 | 2.275 | 5.596 | 0.730 | 0.667 | 1.137 | 3.194 | |
Grid search | 2 | 2.342 | 5.666 | 0.730 | 0.667 | 1.139 | 3.284 |
3 | 2.342 | 5.786 | 0.730 | 0.653 | 1.139 | 3.284 | |
5 | 3.732 | 6.321 | 0.701 | 0.653 | 1.139 | 3.284 | |
Random search | 30 | 3.732 | 6.321 | 0.701 | 0.653 | 2.348 | 3.284 |
40 | 4.765 | 6.432 | 0.692 | 0.664 | 2.348 | 3.284 | |
![]() |
|||||||
Artificial neural network (ANN) | |||||||
Bayesian | 30 | 1.923 | 5.431 | 0.793 | 0.752 | 3.678 | 3.452 |
40 | 1.923 | 5.431 | 0.784 | 0.752 | 2.549 | 4.625 | |
50 | 1.923 | 5.564 | 0.784 | 0.752 | 2.549 | 4.625 | |
Grid search | 2 | 2.987 | 4.320 | 0.784 | 0.764 | 1.493 | 4.625 |
3 | 1.178 | 4.320 | 0.831 | 0.794 | 1.493 | 4.625 | |
5 | 1.178 | 4.320 | 0.835 | 0.799 | 1.493 | 3.356 | |
Random search | 30 | 1.178 | 4.320 | 0.835 | 0.799 | 1.394 | 3.29 |
40 | 1.181 | 4.320 | 0.835 | 0.799 | 1.493 | 3.381 | |
![]() |
|||||||
Support vector machine (SVM) | |||||||
Bayesian | 30 | 2.639 | 8.609 | 0.638 | 0.205 | 1.344 | 4.443 |
Following hyper parameter optimization, the study conducted loss function optimization using three distinct optimizers. For GPR in adsorption capacity, the grid search optimizer with grid sizes of 2, 3, and 5 achieved the lowest error, while for the % removal, the minimum error was attained with grid search using grid sizes of 3 and 5. Using 30 iterations, the random search optimizer yielded a minimal error for ANN in both scenarios. Notably, the SVM model exhibited a significantly high R2 value at the optimized hyper parameters, prompting the decision to forego further optimization for the SVM model. The neural network model with 15 layers, with 128 nodes in the first layer and 256 nodes in the second layer and then the number of nodes increasing in the subsequent layers outperformed all the models in terms of adsorption capacity as well as % removal.
The ANN model (random search) demonstrated better performance than other models, achieving an R2 value of 0.835 for adsorption capacity and 0.799 for % removal. This highlights the ANN model's effectiveness in predicting the percentage of methylene blue removal based on key parameters, like pH, adsorption time, and initial dye concentration. The high R2 values obtained by the ANN and decision tree models provide robust evidence of their accuracy in estimating the adsorption percentage and dye removal capacity. These models can be considered reliable tools for predicting MB adsorption efficiency in diverse scenarios.
This journal is © The Royal Society of Chemistry 2025 |