Environmental water sources are under increasing threat due to the addition of harmful chemicals not addressed by conventional water treatment processes. To work on this concern, the current study aimed to synthesize sono-assisted Fe-modified activated carbon-chitosan (FeAcC) composite and construct a laboratory-scale ozone-integrated fluidized bed reactor (FBR) to eliminate phenolphthalein (php). Following 120 min of incubation, the adsorbent demonstrated a 27.28 mg g−1 of php adsorption capacity at pH 4 with 0.5 g L−1 of adsorbent dosage. The adsorption efficacy and mechanism were defined using isotherm and kinetic models. The study investigated the impact of different factors including, initial concentration, reuse of FeAcC, recirculation flow rate, and hydraulic retention time (HRT), on the efficiency of php removal. The optimum removal efficiency was observed at approximately 95% after 20 min of operation at 1.5 L min−1 recirculation flow rate (batch FBR) and 70 min of HRT (continuous FBR) under 400 mg h−1 ozonation rate. Experimental parameters were optimized using response surface methodology (RSM) with central composite design (CCD) to improve php removal. The large-scale implementation of the findings in the future can be a step for adding new technology for clean water treatment processes for emerging toxic pollutants.

  • Sono-assisted synthesis of Fe-modified activated carbon–chitosan composite for adsorption of phenolphthalein was aimed for in this study.

  • A laboratory-scaled advanced oxidation-integrated fluidized bed reactor was designed.

  • The combination of adsorption and oxidation enhanced the removal rate (≈95%) of phenolphthalein.

  • Response surface methodology was used to experimentally design the optimal experimental condition.

The rapid growth of chemical industries and research laboratories around the globe has led to the increased production and discharge of organic pollutants, posing a substantial risk to the environment. The most concerning compounds are the phenolic compounds (phenol, bisphenol-A (BPA), 4-nonylphenol, chlorophenols, phenolphthalein (php), and octylphenols), which show an adverse impact on all living beings and the environment (Panigrahy et al. 2022). Due to the organoleptic characteristics of phenolic compounds, they are hazardous to aquatic species at concentrations as low as parts per billion. Direct contact and consumption of those compounds may cause liver, eye, skin, muscle, and kidney damage (Udoetok et al. 2016). The most widely utilized phenolic substance is php, disrupting the endocrine system in both humans and animals. It was found that php, an active component of several laxatives, causes cancer in animals across various species (Longnecker et al. 1997). As a result, it is essential to remove organic contaminants efficiently using an affordable technique.

Organic contaminants from aqueous solutions can be removed by a variety of conventional techniques, including biological degradation (Cirik et al. 2013), chemical adsorption (Heydari et al. 2021), electrochemical oxidation (Garciá-Orozco et al. 2016), oxidation (Malik et al. 2019), ion-exchange (Edgar & Boyer 2021), reverse osmosis (Jamil et al. 2020), coagulation, flocculation, and sedimentation (Matilainen et al. 2010). However, these methods might be challenging due to some limitations, including secondary byproducts, being more time-consuming (biological method), high installation costs, improper disposal facility and selective ion-exchange resin (ion-exchange), fouling of membrane and high pressure (reverse osmosis), and toxic chemical utilization (coagulation) (Italiya & Subramanian 2022). Advanced oxidation processes (AOPs) have substantially impacted the effective elimination of harmful phenolic compounds. This is attributed to the generation of hydroxyl radicals (·OH), which oxidize these phenolic compounds (Malik et al. 2019). The adsorption mechanism of the adsorbent material depends on its surface charge and physical properties, which could be significantly altered when it undergoes oxidation (Álvarez et al. 2005). Several published studies have discussed the use of advanced oxidation and adsorption for the elimination of phenolic organic pollutants, such as gallic acid, 17β-estradiol, BPA, phenol, p-chlorophenol, and p-nitrophenol (Álvarez et al. 2005; Irmak et al. 2005; Fan et al. 2019; Lu et al. 2022). The most efficient approach for eliminating organic contaminants is using a fluidized bed reactor (FBR) with enhanced oxidation and adsorption.

The most significant concern in removing toxic php from polluted waterways is to find affordable, flexible, and efficient adsorbent materials. There are numerous adsorbent materials including cobalt–aluminum-layered double oxide (Tongchoo et al. 2020), multi-walled carbon nanotube (Vadi & Namavar 2013), β-cyclodextrin–chitosan (Tang et al. 2019), strontium ferrite (Adewuyi et al. 2021), and tetraethylammonium–kaolinite clay (Adewuyi & Oderinde 2022) that show the adsorptive removal of php. Adsorption of pollutants on chitosan and carbon is limited due to the small surface area and weak mechanical strength. Hence, cross-linking and ultrasound were studied and reported to enhance the cavitation, surface area, and mechanical robustness of the adsorbent material (Italiya et al. 2024). Furthermore, the modification of Fe on chitosan forms a stable composite through complexation and electrostatic interaction. This modification enhances the mechanical durability of chitosan, surface net charge, and hydrophilic–hydrophobic properties alteration and imparts resistance to both acidic and basic conditions (Hu et al. 2016b; Zhang et al. 2018). Thus, phenolic compounds, other organic contaminants, and persistent organic pollutants can be eliminated using modified chitosan and carbon materials (Yazdi et al. 2019). Adsorption with a FBR intimates contact between adsorbate and adsorbent through high mass transfer and rate of reaction with less residence time (Mohapatra et al. 2021). The uniform distribution of adsorbents due to fluidization leads to the adsorption of small quantities of pollutants from an aqueous media (Kalaruban et al. 2016). Adsorption with ozonation and FBR is the most promising substitute method for treating organic pollutants that produce oxygen as a byproduct without producing sludge (Cirik et al. 2013; Italiya & Subramanian 2022).

This research aims to synthesize a sono-assisted Fe-modified activated carbon (Ac)–chitosan (FeAcC) composite to remove php from aqueous solutions. The composite material was characterized using field emission-scanning electron microscopy (FE-SEM), X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), and Brunauer–Emmett–Teller (BET) surface analysis. In the study, the FeAcC demonstrated its efficacy in the batch process for removing php under different conditions, such as pH, initial php concentration, time, and adsorbent dosage. The parameters for isotherm, kinetic, and thermodynamic studies were obtained using the data from the adsorption study. The reactor studies were conducted using an advanced ozone-integrated FBR in the batch and continuous modes of operation. The batch FBR study was carried out with various initial php concentration, flow rate, time, and FeAcC reuse. The continuous FBR study was investigated using the different hydraulic retention times (HRTs). Response surface methodology (RSM) was used to design the optimum experimental condition.

Chemicals

Chitosan (C56H103N9O39, 99.90%), glutaraldehyde ((CH2)3(CHO)2, 98%), and sodium tripolyphosphate (NaTPP, Na5P3O10, 99.99%) were purchased from Sigma Aldrich, India. Phenolphthalein (C20H14O4, 99%), ethanol (CH3CH2OH, 99.99%), charcoal, and ferric chloride hexahydrate (FeCl3·6H2O, 98%) were purchased from SRL Chemical Pvt. Ltd. Sodium hydroxide (NaOH) and hydrochloric acid (HCl) with 0.1M concentration were used to adjust the pH of the solution. The stock solution of php was prepared using an ethanol solution.

Synthesis of sono-assisted FeAcC composite material

For the synthesis of the sono-assisted FeAcC composite, Ac was prepared from commercially available charcoal by pyrolysis. A total weight of 20 g of dried charcoal was measured in a quartz tube and kept inside the muffle furnace at 720 °C (4 °C min−1) under a continuous nitrogen (100 cm3 min−1) supply for 3 h. After reaching the maximum temperature, the nitrogen supply was exchanged with carbon dioxide supply (100 cm3 min−1) for 1 h. Finally, Ac will be collected for the composite preparation (León et al. 2020). Sono-assisted FeAcC was synthesized by dissolving an equal amount of chitosan and Ac into a 2% acetic acid solution. The solution mixture was kept under stirring and sonication conditions for 30 min and 1 h, respectively. To increase the mechanical strength and facilitate the cross-linking of chitosan, the solution mixture was dropwise added to 0.05M NaTPP solution (Yang et al. 2020). The formed beads were subsequently allowed to rest at room temperature for 1 h and washed three times with distilled water to remove excess NaTPP. Following that, beads were stirred for 2 h in a 2.5% glutaraldehyde solution for polymerization. Then beads were washed 3 times and dipped into 0.2M of FeCl3·6H2O solution for 2 h of sonication to impregnate iron on the surface of composite. After that FeAcC beads were rewashed with distilled water 3 times and kept for drying at 60 °C for 24 h. Finally, FeAcC composite material was collected for further adsorption and FBR study. Figure 1 demonstrates the complete methodology of this research work.
Figure 1

Complete methodology of php removal.

Figure 1

Complete methodology of php removal.

Close modal

Characterization of FeAcC composite

FE-SEM was used to study composites' surface structural morphology at various magnifications (Thermo Fisher FEI-Quanta 250 FEG). The crystalline properties of the composite were examined with 2Ɵ range from 10 to 90° using XRD (Bruker D8 advance model). FTIR with a spectral range from 400 to 4,000 cm−1 was used to determine the presence of functional groups on the surface of FeAcC (IRAffinity-1, Shimadzu, Japan). The specific surface area, volume, and size of the pores of the composite were measured under desorption and adsorption of nitrogen gas using a BET surface analyzer (Quantachrome, USA). Before and after treatment, the php concentration was measured using a UV-vis spectrophotometer (Specord 210 plus).

Batch adsorption study

The 20 mg L−1 of php stock solution was prepared to study its adsorption on FeAcC. Since php is known to be poorly soluble in water, ethanol was used to prepare a stock solution and it was diluted using distilled water to adjust the specific concentration. The various parameters including FeAcC dosage (0.1–1 g L−1), pH (3–12), the initial php concentration (5–25 mg L−1), and various reaction times (15–150 min) were followed by a temperature range between 303 and 323 K to determine the optimum condition for php removal. The final concentration of php after adsorption was measured at 552 nm absorbance wavelength by a UV-vis spectrophotometer (Dehabadi et al. 2014). To comprehend the repulsive and attractive mechanism, the FeAcC point zero surface charge (pHzpc) was ascertained using this pH drift method (Italiya et al. 2024). For pHzpc determination, 0.5 g of FeAcC was added to 25 mL of NaCl (0.01M) solution that encompassed a range of pH values (3–12) (Cotman et al. 2016). NaCl was employed as a background electrolyte for the measurement of pHzpc. In aqueous environments, Na+ and Cl ions are rather inert toward the surface of composite materials. The electrolytes' monovalent interaction with FeAcC results in an increase in surface thickness and a limitation on the mobility of H+. That H+ limitation helps in the determination of pHzpc (Kollannur & Arnepalli 2019).

The adsorption percentage and capacity were calculated using Equations (1) and (2), respectively (Karthikeyan et al. 2020):
(1)
(2)
where, is the initial php concentration (mg L−1), is the final php concentration (mg L−1), m is the mass of the composite (g), V is the volume of the working php solution (L), and is the adsorption capacity of FeAcC at equilibrium (mg g−1).

Advanced oxidation-integrated FBR operation

A small lab-scale glass FBR with 200 mL of working volume (3.5 cm internal diameter and 33 cm height) was designed for the php removal study. Figure 2(a) and 2(b) displays the experimental laboratory setup and schematic representation of advanced oxidation-integrated FBR, respectively. The php solution was recirculated in an upward direction (from bottom to top), and FeAcC was continuously fluidized in the reactor through a water pressure pump. To stop the composites from going through the pump during the recirculation process, filters were attached to all entrances of the reactor. The feeding solution was pumped from the top of the reactor using a peristaltic pump (Figure 2(a) and 2(b)). For oxidation, the ozone (400 mg h−1) was supplied from the reactors' bottom by an ozonator purchased from GE Ozone G Pvt. Ltd. All the batch reactor studies were operated in three steps: (i) filling the reactor, (ii) reaction process, and (iii) collecting the solution after treatment. The batch reactor study was carried out with various conditions including initial php concentration (5–25 mg L−1), recirculation flowrate (0.5–2.5 L min−1), adsorbent dose (0.5 g L−1), composite reuse (five runs), and reaction time (2–30 min) to optimize the appropriate removal of php. In the continuous FBR, the reactor was fed with different php concentrations (5–25 mg L−1) at various HRTs (5–200 min), and the maximum removal of php was analyzed. The HRT was maintained by adjusting the inflow rate of feeding solution using a peristaltic pump, and the solution was collected after treatment from the outflow (effluent collection) of a reactor to measure the removal amount of php.
Figure 2

(a) Laboratory-scaled experimental setup of FBR and (b) schematic representation of FBR.

Figure 2

(a) Laboratory-scaled experimental setup of FBR and (b) schematic representation of FBR.

Close modal

Experimental design using RSM

The central composite design (CCD) of the RSM was employed to ascertain the ideal conditions for enhanced removal of php. RSM holds an important advantage in enhancing the statistical significance and interaction of various parameters, which provides better responses for the removal of php. Three crucial independent experimental variables that mostly influenced the php removal are given in Table 1 with their experimental ranges, center points, and levels of independent variables. Design expert 7.0 software was used to evaluate the statistical significance, p-test, F-test, validation, and model prediction of data (Ghadiri et al. 2017). A total of 20 experiments were carried out using 20 runs designed by CCD. The second-order polynomial Equation (3) illustrates how the outcomes of the experimental runs were determined by the removal of php as a response (Salehi & Hosseinifard 2020).
(3)
where Y is the dependent variable (predicted response); , , , and are the constant value, linear effect coefficient, quadratic term, and effect of interactions, respectively; and are the independent variables; and C is the error of prediction.
Table 1

Levels of independent variables and experimental ranges

Sr. No.VariableUnitα−10+1+α
Time min −7.54 16 30 39.54 
Flowrate L min−1 −0.18 0.5 1.5 2.5 3.18 
Initial concentration mg L−1 −1.82 15 25 31.82 
Sr. No.VariableUnitα−10+1+α
Time min −7.54 16 30 39.54 
Flowrate L min−1 −0.18 0.5 1.5 2.5 3.18 
Initial concentration mg L−1 −1.82 15 25 31.82 

Composite characterization

FE-SEM and specific surface area analysis

The morphological structural identification of FeAcC (before and after adsorption of php) is illustrated in Figure 3(a)–3(f) using FE-SEM. The relative size of the spherical FeAcC was between 0.3 and 0.45 mm (Figure 3(a)). The FeAcC composite's porous structure (due to Ac) and uneven surface (due to chitosan) contribute to its increased surface area (Figure 3(b) and 3(c)) (Hu et al. 2016b). The potential adsorption of php was enhanced by the outer uneven surface and inner surface area of pores of FeAcC, which offers a quick and efficient elimination of php from the wastewater and offers more area and contact time between the adsorbate and the adsorbent (Karthikeyan et al. 2019). The density of the FeAcC's surface was high due to the cross-linking reaction between glutaraldehyde and the functional group amine of chitosan to construct a cross-linked network framework by producing a Schiff base (Tang et al. 2019). After adsorption, some of the pores of FeAcC were filled and blocked, which indicates the adsorption and surface complexation of php (Figure 3(d) and 3(e)) (Karthikeyan et al. 2020). Figure 3(b) demonstrates the pleated structure on the surface of FeAcC due to water and ammonia volatilization during moisture removal from the composite (Zhang et al. 2018). The results of SEM analysis support the XRD and FTIR data and exhibit that the synthesis of FeAcC was an efficient adsorbent for the removal of php. The specific surface area and average pore size of FeAcC were determined using BET isotherm and Barret, Joyner, and Halanda adsorption and desorption technique under nitrogen condition, respectively (Figure 3(f)) (Arora et al. 2010). The specific surface area, total pore volume, and mean pore size were 13.67 m2 g−1, 0.10 cm3 g−1, and 2.52 nm, respectively. The small surface area of the FeAcC composite suggested a greater php adsorption capacity.
Figure 3

FE-SEM image of FeAcC composite ((a) magnification: 160 × , (b) magnification: 5,000 × , and (c) magnification: 10,000×) before adsorption and ((d) magnification: 5,000× and (e) magnification: 10,000×) after adsorption; (f) nitrogen adsorption and desorption isotherm.

Figure 3

FE-SEM image of FeAcC composite ((a) magnification: 160 × , (b) magnification: 5,000 × , and (c) magnification: 10,000×) before adsorption and ((d) magnification: 5,000× and (e) magnification: 10,000×) after adsorption; (f) nitrogen adsorption and desorption isotherm.

Close modal

XRD analysis

The crystallographic nature of the FeAcC was examined before and after adsorption of php using XRD with the scan speed of 1° min−1 (Figure 4(a)). The physicochemical alterations in the FeAcC were observed between two theta ranges of 10–90°. The diffraction peaks at 19.90 and 25.27 suggested the semi-crystalline nature of chitosan with their respective characteristic planes (110 and 130) (Aswin Kumar & Viswanathan 2018). The presence of peak values at 29.89, 35.64, 43.1, 53.76, 56.98, and 62.55 indicates a well-distributed and homogeneous spread of Fe on the FeAcC surface, devoid of any contaminants (Mohammadi et al. 2019). Some small peaks between 30–50° demonstrate that the composite's crystalline structure was amorphous and erratically piled by the carbon rings (Bakti & Gareso 2018). The sharp peaks of diffraction (19.90 and 43.10), wide and medium intense peaks (40–60), and extremely wide and less intense peaks suggested the significant size of crystals with the highest degree of crystallization, medium size of crystals with less crystallization degree, and smallest size of crystals with lowest crystallization degree, respectively. The decrease in peak values following the adsorption of php indicated significant structural alterations in the FeAcC (Zhang et al. 2018). The broadening and decreased peak values of 19.90 and 25.27 suggested a reduction or loss of the crystallinity of the composite after adsorption. The peak intensity ranging from 29.89 to 62.55 diminished as a result of the segregation of Fe due to the continuous stirring effect and the attachment of php to the FeAcC surface (Rout et al. 2015; Karthikeyan et al. 2020).
Figure 4

(a) XRD patterns and (b) FTIR spectra of FeAcC before and after adsorption.

Figure 4

(a) XRD patterns and (b) FTIR spectra of FeAcC before and after adsorption.

Close modal

FTIR analysis

FTIR was used to assess the properties of functional groups present in FeAcC. Figure 4(b) provides FTIR spectra of synthesized FeAcC before and after adsorption of php. The strong and broad spectrum near 3,310 cm−1 ascribed the intermolecular hydrogen bonding and vibrational stretching of the amine (N–H) and hydroxyl (–OH) group. The coordination of Fe+3 with hydroxyl and amino groups was suggested by the redshift observed for the FeAcC complex, which resulted in a reduction in hydrogen bonding ability (Eltaweil et al. 2021a). The symmetric and asymmetric vibrational stretching of C–H of methylene and methyl groups are proposed to occur at 2,921 and 2,852 cm−1, respectively (Hu et al. 2016b). The protonation of the amino group of the NHCOCH3 and NH2 was responsible for the appearance of bands near 2,337 and 2,363 cm−1 (Hu et al. 2015). The acetylated amino group of chitosan with C = O vibrational stretching in NH = C = O was represented by an infrared peak at 1,648 cm−1 (Mohammadi et al. 2019). The symmetric C–H vibrational stretching and primary alcohol's stretching vibration C–O are the respective adsorption bands located near 1,384 and 1,085 cm−1. After adsorption, the band at 2,360 cm−1 disappeared as a result of dipole interaction, indicating a reaction between php and protonated amino group of chitosan (Yuan et al. 2008).

Batch adsorption study

Effect of pH on adsorption

The adsorption capacity of the adsorbent material is significantly influenced by the pH of the working solution. The php solution was tested across a pH range of 3–12 to determine the optimal pH for maximum php removal. Alterations in the functional groups and surface charge of both the adsorbent and adsorbate structure have been noted as a result of pH changes (Tahir & Rauf 2006). Slight adsorption of php increased with increasing pH from 3 to 4 and then steadily decreased as the pH increased from 4 to 7. The adsorption capacity increased again between pH 7 and 12. The maximum adsorption capacity was found to be 27.28 mg g−1 at pH 4 (Figure 5(a)). The lower pH value leads to protonation and serves more H+ on the surface of composite material, which is responsible for more H-bond formation between the adsorbate and the adsorbent. Furthermore, the acidic condition imparts a positive charge to the php, which could potentially cause it to repel from the cationic surface of FeAcC (Adewuyi et al. 2021; Shahib et al. 2022). At pH 7, the adsorption capacity was lowest due to neutral surface charge resulting in no electrostatic attraction between php and FeAcC. The php solution becomes negatively charged under an alkaline condition, which leads to electrostatic attraction with the positive amine group of chitosan and surface-modified Fe (Adewuyi & Oderinde 2022). Under alkaline condition, the adsorption capacity was found to be 26.24 mg g−1, which is almost near to the adsorption capacity at pH 4. The positive surface charge of FeAcC was determined based on a pHzpc value of 6.02. The php removal mechanism is depicted in Figure 7.
Figure 5

(a) Effect of pH on adsorption capacity; (b) effect of time and initial concentration on adsorption capacity and (c) percentage removal; (d) effect of adsorbent dosage on adsorption capacity and percentage removal.

Figure 5

(a) Effect of pH on adsorption capacity; (b) effect of time and initial concentration on adsorption capacity and (c) percentage removal; (d) effect of adsorbent dosage on adsorption capacity and percentage removal.

Close modal
Figure 6

Plots of (a) Langmuir, (b) Freundlich, and (c) Temkin isotherms; (d) pseudo-first-order, (e) pseudo-second-order, (f) intraparticle diffusion, and (g) van't Hoff plot thermodynamics.

Figure 6

Plots of (a) Langmuir, (b) Freundlich, and (c) Temkin isotherms; (d) pseudo-first-order, (e) pseudo-second-order, (f) intraparticle diffusion, and (g) van't Hoff plot thermodynamics.

Close modal
Figure 7

Complete php removal mechanism: (a) ozone decomposition, (b) electrostatic attraction, (c) dipole–dipole H-bond, (d) surface complexation, (e) n–π interaction, (f) π–π stacking, and (g) Yoshida's H-bond.

Figure 7

Complete php removal mechanism: (a) ozone decomposition, (b) electrostatic attraction, (c) dipole–dipole H-bond, (d) surface complexation, (e) n–π interaction, (f) π–π stacking, and (g) Yoshida's H-bond.

Close modal

Effect of contact time and initial concentration

The contact time and initial php concentration are important parameters that influence the adsorption study. By varying the contact time from 15 to 150 min and initial concentration from 5 to 25 mg L−1, the sorption of php was investigated. As depicted in Figure 5(b), the removal of php was initially rapid during the beginning of adsorption, but the rate gradually decreased as it approached equilibrium. The adsorption capacity was enhanced by increasing the initial php concentration due to adequate and stable free sites on the FeAcC that interact with low concentrations of php (Yu et al. 2022). The adsorption rate was decreased because of intense competition between php molecules to bind to the surface of FeAcC, which eventually approached saturation by increasing the initial php concentration (Figure 5(c)) (Tang et al. 2019). All of the free active sites were engaged quickly by php at the highest degree of adsorption, which increased the repulsion and reduced the php removal (Adewuyi et al. 2021). The increased mass gradient of php with increasing php concentration works as a driving force to accelerate the php molecules' movement toward the FeAcC (Hu et al. 2016b). The adsorption capacity was found to be 29.46 mg g−1 at the highest initial php concentration (Figure 5(b)). It was found that ≈98% of php was removed for the lowest php concentration at 120 min of reaction time, and after that, there was no drastic change in adsorption (Figure 5(c)). Furthermore, the effect of only ozonation was carried out to remove various concentrations of php (5–25 mg L−1) from an aqueous solution. The maximum removal of php was found to be ≈40% at 45 min for 20 mg L−1 php concentration. There was no significant difference in the removal of php after 20 min of reaction time. The removal of php for all concentrations has been depicted in Supplementary material S1.

Effect of adsorbent dosage

The different FeAcC dosages (0.1–1 g L−1) were tested to determine the ideal quantity of adsorbent for effective php removal (Figure 5(d)). It was observed that the adsorption rate of php was increasing with increasing dose of adsorbent due to the increment of available active sites on the surface of the adsorbent (Eltaweil et al. 2021a). Meanwhile, the adsorption capacity was decreased with increasing the adsorbent dose, because of php aggregation and unsaturation of active sites (Ghadiri et al. 2017). There was a slight increment in the percentage of php adsorption after 0.5 g L−1 of adsorbent dose. Therefore, 0.5 g L−1 adsorbent dose was considered as an optimum dose for the php removal studies.

Adsorption isotherm

Using different concentrations of php ranging from 5 to 25 mg L−1 at pH 4, the adsorption by FeAcC was analyzed at regular intervals to determine the equilibrium point. The residual concentrations of php were examined by collecting the final volume after adsorption at the equilibrium level. Reliable experiment design and the relationship between adsorbate and adsorbent were determined at various temperatures (303–323 K) by adsorption isotherms, including Langmuir, Freundlich, and Temkin models (Mohammadi et al. 2019).

The Langmuir isotherm model provides information about single-layered homogenous adsorption due to the uniform activation energy of the adsorbate (Mohammadi et al. 2019). The maximum adsorption capacity was calculated using vs / plot (Figure 6(a)). The physical and heterogeneous adsorptions are explained by the Freundlich isotherm model. The intensity of adsorption was calculated using the plot (Figure 6(b)) (Tetteh et al. 2020). According to the Temkin model, the repulsion force between the adsorbent and the adsorbate causes the heat of adsorption of each molecule in the multilayer to drop exponentially with the increasing covering layers (Said et al. 2018). The equilibrium binding energy was computed using the plot (Figure 6(c)). The Langmuir, Freundlich, and Temkin isotherms are given below as Equations (4)–(6), respectively:
(4)
(5)
(6)
where, (mg L−1) is the equilibrium concentration of php; (mg g−1) is the efficiency of adsorption at equilibrium; , and are the constants of Langmuir, Freundlich, and Temkin isotherm models, respectively; (mg g−1) is the maximum adsorption capacity; 1/n is the adsorption intensity (0 < 1/n < 1 describes favorable adsorption); is the binding constant at equilibrium concerning highest binding energy; is the heat of adsorption constant; T (K) is the temperature; and is the capability of the adsorbent to uptake the adsorbents.
RL is a constant without dimensions used to describe the key features of the Langmuir isotherm to comprehend the model better; the nonlinear form of RL is expressed in Equation (7) (Meroufel et al. 2013):
(7)
where (mg L−1) is the initial concentration of php solution and RL is the separation factor. According to the literature survey, the values of RL are 0 < RL < 1, RL > 1, RL = 0, and RL = 1 representing favorable, unfavorable, irreversible, and linear adsorption, respectively.

All the values of isotherm parameters are displayed in Table 2. The regression coefficient (R2) values of the Langmuir model at 303, 313, and 323 K are 0.99, 0.99, and 0.99, which show better results in comparison to the Freundlich (0.98, 0.90, and 0.97) and Temkin models (0.96, 0.97, and 0.99), respectively. The results demonstrated that the Langmuir isotherm model was perfectly fit for the php adsorption by FeAcC. The favorable adsorption conditions were represented by the RL values, which ranged from 0 to 1 at different temperatures (Table 2). The Langmuir isotherm parameters indicated that the maximum adsorption capacity () was found to be 25.18 mg g−1 at 323 K. The values of adsorption intensity (1/n) and separation factor (RL) between 0 and 1 suggested the favorable php adsorption condition on the FeAcC surface (Hu et al. 2015; Aswin Kumar & Viswanathan 2018). In this study, it was noted that the adsorption capacity of the FeAcC composite diminished as the temperature increased, which is attributed to the physical adsorption process (Cheah et al. 2013).

Table 2

Parameters of Langmuir, Freundlich, and Temkin isotherms

IsothermsFactorsTemperature (K)
303313323
Langmuir  (mg g−16.28 12.15 25.18 
 5.70 2.73 1.26 
 0.08 0.15 0.28 
 0.99 0.99 0.99 
Freundlich  0.57 0.48 0.34 
 6.22 9.29 13.69 
 0.98 0.90 0.97 
Temkin bT 8.12 7.60 9.14 
 0.49 1.18 2.70 
 0.96 0.97 0.99 
IsothermsFactorsTemperature (K)
303313323
Langmuir  (mg g−16.28 12.15 25.18 
 5.70 2.73 1.26 
 0.08 0.15 0.28 
 0.99 0.99 0.99 
Freundlich  0.57 0.48 0.34 
 6.22 9.29 13.69 
 0.98 0.90 0.97 
Temkin bT 8.12 7.60 9.14 
 0.49 1.18 2.70 
 0.96 0.97 0.99 

Kinetic study

The adsorption efficiency of the FeAcC composite was determined using adsorption kinetic models, including pseudo-first-order, pseudo-second-order, and intraparticle diffusion. The behavior of the php adsorption was studied for 120 min at various temperatures (303, 313, and 323 K) and initial concentrations (5–25 mg L−1). The principle of pseudo-first-order kinetics implies that the variation in equilibrium concentration over time has a direct relationship with the rate of php adsorption (Fegade et al. 2018). According to the pseudo-second-order model, chemisorption is responsible for php adsorption. While intraparticle diffusion has a greater impact on the interaction of adsorbate and surface of adsorbent (Karthikeyan & Meenakshi 2021). The reaction time, mechanism, and relation between time at equilibrium and rate of adsorption can be determined using kinetic models. Pseudo-first-order, pseudo-second-order, and intraparticle diffusion model equations are mentioned below as Equations (8)–(10), respectively:
(8)
(9)
(10)
where (mg g−1) and (mg g−1) are the adsorption of php at equilibrium and at various intervals of time, respectively; (min−1), (g mg−1 min−1), and (mg g−1 min1/2) are the rate constants of pseudo-first-order, pseudo-second-order, and intraparticle diffusion model, respectively; t (min) is the time and h (mg g−1) is the initial rate of adsorption ().

Table 3 lists the different kinetic parameters along with their corresponding R2 and constant values. Figure 6(d)–6(f) represents the plots of pseudo-first-order, pseudo-second-order, and intraparticle diffusion models, respectively. Compared to the pseudo-first-order kinetics and intraparticle diffusion models, the pseudo-second-order model fits the data most accurately, as shown by their R2 values. Furthermore, the calculated values appear to be almost close to the experimental values, suggesting the suitability of a pseudo-second-order model for php adsorption. The chemisorption mechanism of php adsorption on the surface of FeAcC was suggested by the decreasing values with increasing initial php concentration (Eltaweil et al. 2021b). The occurrence of php adsorption during chemisorption is facilitated by the exchange of electrons under the influence of valence forces (Cheah et al. 2013). Based on the kinetic study, the php adsorption rate was dependent upon available reactive binding sites and porosity of the FeAcC (Italiya et al. 2022).

Table 3

Kinetic parameters for adsorption of php by FeAcC

(mg L−1)510152025
(mg g−19.25 16.97 22.69 25.48 29.47 
Pseudo-first-order  (mg g−19.16 14.83 16.76 22.24 26.60 
 2.43 × 10−4 2.34 × 10−4 4.02 × 10−4 2.50 × 10−4 2.20 × 10−4 
R2 0.70 0.71 0.51 0.81 0.96 
Pseudo-second-order  (mg g−111.30 16.45 21.59 25.65 28.12 
 2.10 × 10−3 1.07 × 10−3 1.40 × 10−3 9.11 × 10−4 5.29 × 10−4 
R2 0.97 0.98 0.96 0.95 0.98 
Intraparticle diffusion  0.74 1.39 1.55 1.99 1.88 
R2 0.90 0.88 0.94 0.95 0.98 
(mg L−1)510152025
(mg g−19.25 16.97 22.69 25.48 29.47 
Pseudo-first-order  (mg g−19.16 14.83 16.76 22.24 26.60 
 2.43 × 10−4 2.34 × 10−4 4.02 × 10−4 2.50 × 10−4 2.20 × 10−4 
R2 0.70 0.71 0.51 0.81 0.96 
Pseudo-second-order  (mg g−111.30 16.45 21.59 25.65 28.12 
 2.10 × 10−3 1.07 × 10−3 1.40 × 10−3 9.11 × 10−4 5.29 × 10−4 
R2 0.97 0.98 0.96 0.95 0.98 
Intraparticle diffusion  0.74 1.39 1.55 1.99 1.88 
R2 0.90 0.88 0.94 0.95 0.98 

Thermodynamic study

The temperature effect on the rate of php adsorption was investigated by important thermodynamic parameters, including standard entropy change (Δ) and Gibbs free energy change (Δ). Equations (11) and (12) were used to calculate Δ and Δ, respectively.
(11)
(12)
where is the thermodynamic equilibrium constant (L g−1), R is the global gas constant (J mol−1 K−1), and T is the temperature (K).

The R2 value (0.99) was calculated using the Van't Hoff plot (ln k vs 1/T) (Figure 6(g)). Adsorption appeared to be spontaneous and thermodynamically feasible, as indicated by the negative ΔG° (Supplementary material S2). Furthermore, the high temperature was shown to be more favorable for the adsorption of php, as seen by a rise in negative ΔG° with temperature. The increasing randomness at the solution/solid interface following the adsorption of php was confirmed by the positive value of ΔS° (46.96 kJ K−1 mol−1). The positive value of ΔH° (11.78 kJ mol−1) indicated the endothermic nature of the php adsorption process.

Comparison study

Table 4(a) and 4(b) display the comparative adsorption capacity of FeAcC with various php initial concentrations and removal of other organic compounds using an ozone-integrated reactor study, respectively. Concerning the small dose and surface area of FeAcC and low ozonation rate, the php removal capacity and rate were high in comparison to other composites listed in Table 4.

Table 4

Comparative study of (a) php adsorption capacity with synthesized FeAcC and (b) other organic pollutant removal study using ozone-integrated reactor study

(a) Removal of php by adsorption mechanism
Sr. No.Adsorbent materialAdsorption capacityReferences
Cobalt/aluminum-layered double oxide 22.6 Tongchoo et al. (2020)  
Cobalt/aluminum-layered double hydroxide 8.0   
Tetraethylammonium/kaolinite clay 43 Adewuyi & Oderinde (2022)  
Supramolecular recognized polyvinylidene fluoride/polyvinyl alcohol – Lu et al. (2022)  
β-cyclodextrin/polystyrene fibers 7.09 Uyar et al. (2009)  
Strontium ferrite 7.40 Adewuyi et al. (2021)  
Polyvinylidene difluoride – Guo et al. (2023)  
β-cyclodextrin/chitosan 0.724 Tang et al. (2019)  
β-cyclodextrin/carbon/cobalt nanomagnet 2.0 Fuhrer et al. (2011)  
10 β-cyclodextrin/polyurethanes 3.97 Mohamed et al. (2011)  
11 Polyvinylidene fluoride/polystyrene/polydopamine-24/β-cyclodextrin 19.24 Gao et al. (2019)  
12 Polyvinylidene fluoride/polystyrene/polydopamine-12/β-cyclodextrin 15.23   
13 β-cyclodextrin/polyurethanes (1,6-hexamethylene diisocyanate) 19.2 Mohamed & Wilson (2015)  
14 β-cyclodextrin/polyurethanes (4,4-dicyclohexylmethane diisocyanate) 3.13   
15 β-cyclodextrin/polyurethanes (4,4-diphenylmethane diisocyanate) 14.7   
16 β-cyclodextrin/polyurethanes (1,4-phenylene diisocyanate) 2.17   
17 FeAcC 25.18 Present study 
(a) Removal of php by adsorption mechanism
Sr. No.Adsorbent materialAdsorption capacityReferences
Cobalt/aluminum-layered double oxide 22.6 Tongchoo et al. (2020)  
Cobalt/aluminum-layered double hydroxide 8.0   
Tetraethylammonium/kaolinite clay 43 Adewuyi & Oderinde (2022)  
Supramolecular recognized polyvinylidene fluoride/polyvinyl alcohol – Lu et al. (2022)  
β-cyclodextrin/polystyrene fibers 7.09 Uyar et al. (2009)  
Strontium ferrite 7.40 Adewuyi et al. (2021)  
Polyvinylidene difluoride – Guo et al. (2023)  
β-cyclodextrin/chitosan 0.724 Tang et al. (2019)  
β-cyclodextrin/carbon/cobalt nanomagnet 2.0 Fuhrer et al. (2011)  
10 β-cyclodextrin/polyurethanes 3.97 Mohamed et al. (2011)  
11 Polyvinylidene fluoride/polystyrene/polydopamine-24/β-cyclodextrin 19.24 Gao et al. (2019)  
12 Polyvinylidene fluoride/polystyrene/polydopamine-12/β-cyclodextrin 15.23   
13 β-cyclodextrin/polyurethanes (1,6-hexamethylene diisocyanate) 19.2 Mohamed & Wilson (2015)  
14 β-cyclodextrin/polyurethanes (4,4-dicyclohexylmethane diisocyanate) 3.13   
15 β-cyclodextrin/polyurethanes (4,4-diphenylmethane diisocyanate) 14.7   
16 β-cyclodextrin/polyurethanes (1,4-phenylene diisocyanate) 2.17   
17 FeAcC 25.18 Present study 
(b) Removal of organic compounds by ozonation and reactor study
Sr. No.CatalystPollutantOzone dose% of degradationReferences
Granular activated carbon Textile wastewater 4 L min−1 93 decolorization Lin & Lai (2000)  
Carvon aerogel/copper oxide Textile dye wastewater 4 mg min−1 85 decolorization Hu et al. (2016a)  
Manganese/copper- Al2O3 Benzotriazole 2.6 g h−1 77 mineralization Roshani et al. (2014)  
Activated carbon Phenol 3 L min−1 51.53 Xiong et al. (2020)  
TiO2 Diclofenac and amoxicillin 50 mg L−1 Almost 100 Moreira et al. (2015)  
Activated carbon Azo dye acid red 27 0.80 mg h−1 ≈ 90 Beak et al. (2009)  
Activated carbon fiber Phenolic wastewater 1.5 L min−1 99 Qu et al. (2007)  
Ruthenium/γ-Al2O3 BPA 6 L h−1 56 mineralization Cotman et al. (2016)  
Granular activated carbon Phenolic compounds 19 mg L−1 94 mineralization Ferreiro et al. (2021)  
10 FeAcC php 400 mg h−1 ≈ 95 Present study 
(b) Removal of organic compounds by ozonation and reactor study
Sr. No.CatalystPollutantOzone dose% of degradationReferences
Granular activated carbon Textile wastewater 4 L min−1 93 decolorization Lin & Lai (2000)  
Carvon aerogel/copper oxide Textile dye wastewater 4 mg min−1 85 decolorization Hu et al. (2016a)  
Manganese/copper- Al2O3 Benzotriazole 2.6 g h−1 77 mineralization Roshani et al. (2014)  
Activated carbon Phenol 3 L min−1 51.53 Xiong et al. (2020)  
TiO2 Diclofenac and amoxicillin 50 mg L−1 Almost 100 Moreira et al. (2015)  
Activated carbon Azo dye acid red 27 0.80 mg h−1 ≈ 90 Beak et al. (2009)  
Activated carbon fiber Phenolic wastewater 1.5 L min−1 99 Qu et al. (2007)  
Ruthenium/γ-Al2O3 BPA 6 L h−1 56 mineralization Cotman et al. (2016)  
Granular activated carbon Phenolic compounds 19 mg L−1 94 mineralization Ferreiro et al. (2021)  
10 FeAcC php 400 mg h−1 ≈ 95 Present study 

Batch FBR study

Effect of initial php concentration and php removal mechanism

Effects of various initial php concentrations (5–25 mg L−1) were examined in an ozone-supplied batch reactor study at a pH of 10. Because the alkaline condition provides more –OH groups to interact with O3 and production of free radicals (·OH), increased concentration of php will logically limit the rate of removal. It was observed that maximum removal occurred during the initial 20 min, and after that, there was no drastic change in php removal. The removal rate of php decreased from 99.02 to 95.60% with an increased initial concentration from 5 to 25 mg L−1 (Figure 8(a)). For the 20 mg L−1 concentration, the removal percentage increased from 63.71% (Figure 5(c)) to 94.19% (Figure 8(a)) and the reaction time decreased from 120 to 20 min after the ozone-supplied batch FBR study. Almost ≈30% of php removal was examined due to combined adsorption and oxidation effects under fluidized conditions. When compared to batch adsorption studies, it was demonstrated that the use of an ozone-supplied batch FBR study improved the removal of php with increasing concentration. The increased rate of php removal occurred due to the production of ·OH via self-decomposition of ozone (Equations (13)–(17)), an ozone transformation (Equations (18)–(21)), and interaction of ozone with Fe (Equations (22)–(25)) (Kasprzyk-Hordern et al. 2003; Malik et al. 2020). The alkaline condition provides more hydroxyl groups to react with ozone and increases the level of ·OH production, which enhances the rate of reaction. Ac helps in more production of ·OH by transforming the ozone throughout catalysis (KISHIMOTO & UENO 2012). The FeAcC composite material provides a surface to adsorb and dissolve the maximum amount of ozone, which leads to ozone decomposition (Figure 7(a)) (Li et al. 2015). The alkaline condition makes the php more negative, which increases the php's attraction to the positively charged FeAcC through ion-exchange and electrostatic interaction (Figure 7(b)) (Adewuyi & Oderinde 2022). The slightly positive hydrogen atom of FeAcC interacts with the slightly negative oxygen atom of php and forms dipole–dipole H-bonds to eliminate php from the aqueous solution (Figure 7(c)) (Mohamed et al. 2011). The corresponding php anions interact with the surface metal ions through an inner or outer surface mechanism to produce a surface complexation (Figure 7(d)) (Karthikeyan et al. 2020). In the adsorption mechanism, n–π interaction (between the carbonyl oxygen group of FeAcC and aromatic ring of php) (Figure 7(e)) and π–π interaction (between both the aromatic rings of php and FeAcC) (Figure 7(f)) play an important role (Tran et al. 2017; Shahib et al. 2022). Yoshida's H-bond formed between the aromatic rings of php and the –OH group of FeAcC, allowing php to be adsorbed on the surface of FeAcC (Figure 7) (Tran et al. 2017). Figure 7 represents the complete php removal mechanism by physicochemical adsorption and AOPs.
(13)
(14)
(15)
(16)
(17)
(18)
(19)
(20)
(21)
(22)
(23)
(24)
(25)
Figure 8

Effect of (a) various php concentration, (b) recirculation flowrate, (c) FeAcC reuse on php removal in batch FBR, and (d) HRT on various php concentration in continuous FBR study.

Figure 8

Effect of (a) various php concentration, (b) recirculation flowrate, (c) FeAcC reuse on php removal in batch FBR, and (d) HRT on various php concentration in continuous FBR study.

Close modal

Effect of recirculation flowrate

Figure 8(b) shows the effect of various recirculation flow rates (0.5–2.5 L min−1) on the php removal. The percentage removal of php increased with the increasing flow rate up to 1.5 L min−1. When the flow rate was low, there was a longer duration for php to engage with FeAcC. However, the interaction was confined to the lower section of the reactor because of the extremely slow movement. At the 1.5 L min−1 flow rate, the removal efficiency was highest due to the distribution of php throughout the reactor with maximum contact with the adsorbent and ozone to get adsorbed and oxidized, respectively. Beyond the flow rate of 1.5 L min−1, the removal rate of php was decreased due to the low residence time of •OH to react and leaching of Fe by the high fluid force. Therefore, it was essential to keep the recirculation flow rate at 1.5 L min−1 to get maximum removal of php (Mohapatra & Ghosh 2023).

Effect of FeAcC reusability

The recovery and subsequent utilization of FeAcC offers a cost-effective approach for php removal. The FeAcC reusability experiment was carried out in five cycles (runs) using an optimum dosage of composite (0.5 g L−1) and recirculation flow rate (1.5 L min−1). The FBR introduced the php solution, which was treated with the HRT of 30 min and disposed of at the end of each cycle. Cycle 1 used FeAcC material, which was reused for all the cycles to remove the subsequent php solution. Figure 8(c) shows that the removal rate of php decreased with an increasing number of cycles from 94.74% in run 1 to 56.84% in run 5. This is most likely due to the occupied binding active sites during each cycle (Li et al. 2015). With increasing cycles, repulsion between adsorbed and non-adsorbed php reduces the php uptake rate. The decrease in php removal could also be attributed to the reduced ozone adsorption and decomposition, which may be due to the limited availability of active sites and Fe leaching after multiple uses of FeAcC (Mohapatra & Ghosh 2023).

Continuous FBR study

Effect of HRT

For the continuous FBR study, HRT is a crucial parameter, representing the average duration that php remains within the reactor. To identify the optimal parameters for php removal, the experiment was conducted with varying initial concentrations of php at different HRTs (Figure 8(d)). The feed rate of php affects HRT, and a low feed rate results in high HRT and vice versa. The removal rate was low at the low HRT due to the high feeding rate of php, less time to interact with FeAcC and •OH, and high dragging force generated by the collision between php solution and FeAcC. At high HRT, php had more time to interact with FeAcC and ·OH without any collision. The removal rate decreased with increased concentration of php (Mohapatra et al. 2021). The optimum HRT was found to be 70 min with the highest php removal for all the php concentrations (Figure 8(d)). According to reports, ideal HRT relies on the type of reactor, fluidized material, pollutant, and feeding rate (Sinharoy et al. 2019).

Optimization of php removal using CCD

The relationship between the dependent response and independent factors as well as the calculation of the php elimination efficiency were demonstrated using the quadratic second-order polynomial Equation (26).
(26)
where Y is the php removal percentage and A, B, and C represent the time (min), flow rate (L min−1), and initial php concentration (mg L−1), respectively.
The predicted model showed a significant fit, as evidenced by the p-value <0.05 and F-value of 452.07 shown in Supplementary material S3. The R2 value (0.99) of the developed model represents the valid and appropriate fitting of experimental data. The model goodness and accuracy were determined by the predicted R2. The adjusted R2 demonstrated the correlation between the fitted model and experimental data. The suitability of the developed model was offered by the and R2 of the model. The models' sufficient signal of precision and reliability were determined by calculating the ratio of adequate precision (64.39 > 4) (Salehi & Hosseinifard 2020). The lack of fit (1.53) was considered nonsignificant compared to the pure error. According to the RSM study, the maximum php removal study was observed at 1.5 L min−1 flow rate, 5 mg L−1 php concentration, and 16 min time. The constructed model was confirmed to be accurate by the linear relationship between studentized residuals and normal probability (Figure 9(a)), as well as between actual and predicted values (Figure 9(b)) (Nejatbakhsh et al. 2023). Figure 9(c)–9(e) represents the combined effect of time and flow rate, initial concentration and flow rate, and initial concentration and time against the php removal, respectively. The combined effect plots of Figure 9 along with models' F- value, p-value, and R2 value indicate that each variable (time, flow rate, and initial concentration) is crucial to remove php from an aqueous solution (Italiya et al. 2022).
Figure 9

The parity plots of (a) normal probability vs studentized residues and (b) predicted vs actual values; response surface 3D plot of percentage of php removal vs (c) flow rate and time, (d) initial concentration and flow rate, and (e) initial concentration and time.

Figure 9

The parity plots of (a) normal probability vs studentized residues and (b) predicted vs actual values; response surface 3D plot of percentage of php removal vs (c) flow rate and time, (d) initial concentration and flow rate, and (e) initial concentration and time.

Close modal

In this study, a sono-assisted synthesized FeAcC composite material was used for the removal of php from an aqueous solution. The maximum adsorption capacity was found to be 25.18 mg g−1 using the Langmuir adsorption isotherm. The electrostatic mechanism played an important role during the adsorption process. According to the Langmuir isotherm and pseudo-second-order kinetics models, adsorption occurred in a single homogenous layer with a chemisorption mechanism. Ozone-integrated FBR showed around 30% more php removal within 20 min than a single adsorption technique. The Ac and Fe of FeAcC provided greater surface area to react with ozone to produce more •OH radicals for php removal. A combined adsorptive–oxidative removal of php in FBR enhanced the removal efficiency, residence time, removal at the highest and lowest concentration, no byproducts, and high rate of reaction by resolving the drawbacks of each method when employed independently. The RSM study demonstrated the reliability and significance of the developed model from the experimental data to optimize the parameters. Future research must use real-time wastewater as the study topic to understand better the removal of other organic pollutants by fluidized FeAcC using advanced oxidation-integrated FBR.

The authors acknowledge Vellore Institute of Technology, Vellore, Tamil Nadu, for providing a better opportunity to carry out their studies.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Adewuyi
A.
&
Oderinde
R. A.
2022
Removal of phenolphthalein and methyl orange from laboratory wastewater using tetraethylammonium modified kaolinite clay
.
Current Research in Green and Sustainable Chemistry
5
(
5
),
1
13
.
https://doi.org/10.1016/j.crgsc.2022.100320
.
Adewuyi
A.
,
Gervasi
C. A.
&
Mirífico
M. V.
2021
Synthesis of strontium ferrite and its role in the removal of methyl orange, phenolphthalein and bromothymol blue from laboratory wastewater
.
Surfaces and Interfaces
27
(
9
),
1
16
.
https://doi.org/10.1016/j.surfin.2021.101567
.
Álvarez
P. M.
,
García-Araya
J. F.
,
Beltrán
F. J.
,
Masa
F. J.
&
Medina
F.
2005
Ozonation of activated carbons: Effect on the adsorption of selected phenolic compounds from aqueous solutions
.
Journal of Colloid and Interface Science
283
(
2
),
503
512
.
http://dx.doi.org/10.1016/j.jcis.2004.09.014
.
Arora
M.
,
Eddy
N. K.
,
Mumford
K. A.
,
Baba
Y.
,
Perera
J. M.
&
Stevens
G. W.
2010
Surface modification of natural zeolite by chitosan and its use for nitrate removal in cold regions
.
Cold Regions Science and Technology
62
(
2–3
),
92
97
.
http://dx.doi.org/10.1016/j.coldregions.2010.03.002
.
Aswin Kumar
I.
&
Viswanathan
N.
2018
Development and reuse of amine-grafted chitosan hybrid beads in the retention of nitrate and phosphate
.
Journal of Chemical and Engineering Data
63
(
1
),
147
158
.
10.1021/acs.jced.7b00751
.
Bakti
A. I.
&
Gareso
P. L.
2018
Characterization of active carbon prepared from coconuts shells using FTIR, XRD and SEM techniques
.
Jurnal Ilmiah Pendidikan Fisika Al-Biruni
7
(
1
),
33
39
.
https://doi.org/10.24042/jipfalbiruni.v7i1.2459
.
Beak
M. H.
,
Ijagbemi
C. O.
&
Kim
D. S.
2009
Azo dye acid red 27 decomposition kinetics during ozone oxidation and adsorption processes
.
Journal of Environmental Science and Health – Part A Toxic/Hazardous Substances and Environmental Engineering
44
(
6
),
623
629
.
https://doi.org/10.1080/10934520902784708
.
Cheah
W.
,
Hosseini
S.
,
Khan
M. A.
,
Chuah
T. G.
&
Choong
T. S. Y.
2013
Acid modified carbon coated monolith for methyl orange adsorption
.
Chemical Engineering Journal
215–216
(
7
),
747
754
.
http://dx.doi.org/10.1016/j.cej.2012.07.004
.
Cirik
K.
,
Dursun
N.
,
Sahinkaya
E.
&
Çinar
Ö.
2013
Effect of electron donor source on the treatment of Cr(VI)-containing textile wastewater using sulfate-reducing fluidized bed reactors (FBRs)
.
Bioresource Technology
133
(
1
),
414
420
.
http://dx.doi.org/10.1016/j.biortech.2013.01.064
.
Cotman
M.
,
Erjavec
B.
,
Djinović
P.
&
Pintar
A.
2016
Catalyst support materials for prominent mineralization of bisphenol A in catalytic ozonation process
.
Environmental Science and Pollution Research
23
(
10
),
10223
10233
.
http://dx.doi.org/10.1007/s11356-016-6251-y
.
Dehabadi
V. A.
,
Buschmann
H. J.
&
Gutmann
J. S.
2014
Spectrophotometric estimation of the accessible inclusion sites of β-cyclodextrin fixed on cotton fabrics using phenolic dyestuffs
.
Analytical Methods
6
(
10
),
3382
3387
.
http://dx.doi.org/10.1039/c4ay00293h
.
Edgar
M.
&
Boyer
T. H.
2021
Removal of natural organic matter by ion exchange: Comparing regenerated and non-regenerated columns
.
Water Research
189
(
11
),
1
12
.
https://doi.org/10.1016/j.watres.2020.116661
.
Eltaweil
A. S.
,
Abd
E. M.
,
Monaem
E.
,
Eldin
M. S. M.
&
Omer
A. M.
2021a
Fabrication of attapulgite/magnetic aminated chitosan composite as efficient and reusable adsorbent for Cr (VI) ions
.
Scientific Reports
189
(
11
),
1
15
.
https://doi.org/10.1038/s41598-021-96145-6
.
Eltaweil
A. S.
,
Omer
A. M.
,
El-Aqapa
H. G.
,
Gaber
N. M.
,
Attia
N. F.
,
El-Subruiti
G. M.
,
Mohy-Eldin
M. S.
&
Abd El-Monaem
E. M.
2021b
Chitosan based adsorbents for the removal of phosphate and nitrate: A critical review
.
Carbohydrate Polymers
274
(
9
),
1
16
.
https://doi.org/10.1016/j.carbpol.2021.118671
.
Fan
S.
,
Huang
Z.
,
Zhang
Y.
,
Hu
H.
,
Liang
X.
,
Gong
S.
,
Zhou
J.
&
Tu
R.
2019
Magnetic chitosan-hydroxyapatite composite microspheres: Preparation, characterization, and application for the adsorption of phenolic substances
.
Bioresource Technology
274
(
10
),
48
55
.
https://doi.org/10.1016/j.biortech.2018.11.078
.
Fegade
U.
,
Jethave
G.
,
Su
K. Y.
,
Huang
W. R.
&
Wu
R. J.
2018
An multifunction Zn0.3Mn0.4O4 nanospheres for carbon dioxide reduction to methane via photocatalysis and reused after five cycles for phosphate adsorption
.
Journal of Environmental Chemical Engineering
6
(
2
),
1918
1925
.
https://doi.org/10.1016/j.jece.2018.02.040
.
Ferreiro
C.
,
de Luis
A.
,
Villota
N.
,
Lomas
J. M.
,
Lombraña
J. I.
&
Camarero
L. M.
2021
Application of a combined adsorption-ozonation process for phenolic wastewater treatment in a continuous fixed-bed reactor
.
Catalysts
11
(
8
),
1
20
.
https://doi.org/10.3390/catal11081014
.
Fuhrer
R.
,
Herrmann
I. K.
,
Athanassiou
E. K.
,
Grass
R. N.
&
Stark
W. J.
2011
Immobilized β-cyclodextrin on surface-modified carbon-coated cobalt nanomagnets: Reversible organic contaminant adsorption and enrichment from water
.
Langmuir
27
(
5
),
1924
1929
.
https://doi.org/10.1021/la103873v
.
Gao
N.
,
Yang
J.
,
Wu
Y.
,
Yue
J.
,
Cao
G.
,
Zhang
A.
,
Ye
L.
&
Feng
Z.
2019
β-Cyclodextrin functionalized coaxially electrospun poly(vinylidene fluoride) @ polystyrene membranes with higher mechanical performance for efficient removal of phenolphthalein
.
Reactive and Functional Polymers
141
(
4
),
100
111
.
https://doi.org/10.1016/j.reactfunctpolym.2019.05.001
.
Garciá-Orozco
V. M.
,
Barrera-Diáz
C. E.
,
Roa-Morales
G.
&
Linares-Hernández
I.
2016
A comparative electrochemical-ozone treatment for removal of phenolphthalein
.
Journal of Chemistry
2016
(
9
),
1
9
.
https://doi.org/10.1155/2016/8105128
.
Ghadiri
S. K.
,
Nasseri
S.
,
Nabizadeh
R.
,
Khoobi
M.
,
Nazmara
S.
&
Mahvi
A. H.
2017
Adsorption of nitrate onto anionic bio-graphene nanosheet from aqueous solutions: Isotherm and kinetic study
.
Journal of Molecular Liquids
242
(
6
),
1111
1117
.
http://dx.doi.org/10.1016/j.molliq.2017.06.122
.
Guo
D.
,
Zhang
X.
,
Xie
J.
&
Li
N.
2023
Effect of cross-linking agent concentration on adsorption performance of PVDF ultrafiltration membrane
.
European Polymer Journal
198
(
7
),
1
9
.
https://doi.org/10.1016/j.eurpolymj.2023.112332
.
Heydari
S.
,
Zare
L.
&
Eshagh Ahmadi
S.
2021
Removal of phenolphthalein by aspartame functionalized dialdehyde starch nano-composite and optimization by Plackett–Burman design
.
Journal of the Iranian Chemical Society
18
(
12
),
3417
3427
.
https://doi.org/10.1007/s13738-021-02275-z
.
Hu
Q.
,
Chen
N.
,
Feng
C.
&
Hu
W. W.
2015
Nitrate adsorption from aqueous solution using granular chitosan-Fe 3+ complex
.
Applied Surface Science
347
(
4
),
1
9
.
http://dx.doi.org/10.1016/j.apsusc.2015.04.049
.
Hu
E.
,
Wu
X.
,
Shang
S.
,
Tao
X. M.
,
Jiang
S. X.
&
Gan
L.
2016a
Catalytic ozonation of simulated textile dyeing wastewater using mesoporous carbon aerogel supported copper oxide catalyst
.
Journal of Cleaner Production
112
(
6
),
4710
4718
.
http://dx.doi.org/10.1016/j.jclepro.2015.06.127
.
Hu
Q.
,
Chen
N.
,
Feng
C.
,
Hu
W.
,
Zhang
J.
,
Liu
H.
&
He
Q.
2016b
Nitrate removal from aqueous solution using granular chitosan-Fe(III)-Al(III) complex: Kinetic, isotherm and regeneration studies
.
Journal of the Taiwan Institute of Chemical Engineers
63
(
3
),
216
225
.
https://doi.org/10.1016/j.jtice.2016.03.004
.
Irmak
S.
,
Erbatur
O.
&
Akgerman
A.
2005
Degradation of 17β-estradiol and bisphenol A in aqueous medium by using ozone and ozone/UV techniques
.
Journal of Hazardous Materials
126
(
5
),
54
62
.
https://doi.org/10.1016/j.jhazmat.2005.05.045
.
Italiya
G.
,
Ahmed
M. H.
&
Subramanian
S.
2022
Titanium oxide bonded zeolite and bentonite composites for adsorptive removal of phosphate
.
Environmental Nanotechnology, Monitoring and Management
17
(
12
),
1
16
.
https://doi.org/10.1016/j.enmm.2022.100649
.
Italiya
G.
,
Dongaonkar
A.
&
Jayakumar
N.
2024
Sono-assisted adsorption of nitrate by Fe-modified chitosan–zeolite composite material
.
Journal of Dispersion Science and Technology
2
(
3
),
1
12
.
https://doi.org/10.1080/01932691.2024.2325390
.
Jamil
S.
,
Loganathan
P.
,
Kandasamy
J.
,
Listowski
A.
,
McDonald
J. A.
,
Khan
S. J.
&
Vigneswaran
S.
2020
Removal of organic matter from wastewater reverse osmosis concentrate using granular activated carbon and anion exchange resin adsorbent columns in sequence
.
Chemosphere
261
(
7
),
1
9
.
https://doi.org/10.1016/j.chemosphere.2020.127549
.
Kalaruban
M.
,
Loganathan
P.
,
Shim
W. G.
,
Kandasamy
J.
,
Naidu
G.
,
Nguyen
T. V.
&
Vigneswaran
S.
2016
Removing nitrate from water using iron-modified Dowex 21K XLT ion exchange resin: Batch and fluidised-bed adsorption studies
.
Separation and Purification Technology
158
(
12
),
62
70
.
http://dx.doi.org/10.1016/j.seppur.2015.12.022
.
Karthikeyan
P.
&
Meenakshi
S.
2021
Fabrication of hybrid chitosan encapsulated magnetic-kaolin beads for adsorption of phosphate and nitrate ions from aqueous solutions
.
International Journal of Biological Macromolecules
168
(
11
),
750
759
.
https://doi.org/10.1016/j.ijbiomac.2020.11.132
.
Karthikeyan
P.
,
Banu
H. A. T.
&
Meenakshi
S.
2019
Synthesis and characterization of metal loaded chitosan-alginate biopolymeric hybrid beads for the efficient removal of phosphate and nitrate ions from aqueous solution
.
International Journal of Biological Macromolecules
2
(
2
),
407
418
.
https://doi.org/10.1016/j.ijbiomac.2019.02.059
.
Karthikeyan
P.
,
Vigneshwaran
S.
&
Meenakshi
S.
2020
Al3+ incorporated chitosan-gelatin hybrid microspheres and their use for toxic ions removal: Assessment of its sustainability metrics
.
Environmental Chemistry and Ecotoxicology
2
(
7
),
97
106
.
https://doi.org/10.1016/j.enceco.2020.07.004
.
Kasprzyk-Hordern
B.
,
Ziółek
M.
&
Nawrocki
J.
2003
Catalytic ozonation and methods of enhancing molecular ozone reactions in water treatment
.
Applied Catalysis B: Environmental
46
(
4
),
639
669
.
https://doi.org/10.1016/S0926-3373(03)00326-6
.
Kishimoto
N.
&
Ueno
S.
2012
Catalytic effect of several iron species on ozonation
.
Journal of Water and Environment Technology
10
(
2
),
205
215
.
https://doi.org/10.2965/jwet.2012.205
.
Kollannur
N. J.
&
Arnepalli
D. N.
2019
Methodology for determining point of zero salt effect of clays in terms of surface charge properties
.
Journal of Materials in Civil Engineering
31
(
12
),
1
9
.
https://doi.org/10.1061/(ASCE)MT.1943-5533.0002947
.
León
M.
,
Silva
J.
,
Carrasco
S.
&
Barrientos
N.
2020
Design, cost estimation and sensitivity analysis for a production process of activated carbon from waste nutshells by physical activation
.
Processes
8
(
945
),
1
14
.
https://doi.org/10.3390/pr8080945
.
Li
H.
,
Priambodo
R.
,
Wang
Y.
,
Zhang
H.
&
Huang
Y. H.
2015
Mineralization of bisphenol A by photo-Fenton-like process using a waste iron oxide catalyst in a three-phase fluidized bed reactor
.
Journal of the Taiwan Institute of Chemical Engineers
53
(
2
),
68
73
.
http://dx.doi.org/10.1016/j.jtice.2015.02.024
.
Lin
S. H.
&
Lai
C. L.
2000
Kinetic characteristics of textile wastewater ozonation in fluidized and fixed activated carbon beds
.
Water Research
34
(
3
),
763
772
.
https://doi.org/10.1016/S0043-1354(99)00214-6
.
Longnecker
M. P.
,
Sandler
D. P.
,
Haile
R. W.
&
Sandler
R. S.
1997
Phenolphthalein-containing laxative use in relation to adenomatous colorectal polyps in three studies
.
Environmental Health Perspectives
105
(
11
),
1210
1212
.
https://doi.org/10.1289/ehp.971051210
.
Lu
Q.
,
Li
N.
&
Zhang
X.
2022
Supramolecular recognition PVDF/PVA ultrafiltration membrane for rapid removing aromatic compounds from water
.
Chemical Engineering Journal
436
(
7
),
1
15
.
https://doi.org/10.1016/j.cej.2021.132889
.
Malik
S. N.
,
Khan
S. M.
,
Ghosh
P. C.
,
Vaidya
A. N.
,
Das
S.
&
Mudliar
S. N.
2019
Nano catalytic ozonation of biomethanated distillery wastewater for biodegradability enhancement, color and toxicity reduction with biofuel production
.
Chemosphere
230
(
1
),
449
461
.
https://doi.org/10.1016/j.chemosphere.2019.05.067
.
Malik
S. N.
,
Ghosh
P. C.
,
Vaidya
A. N.
&
Mudliar
S. N.
2020
Hybrid ozonation process for industrial wastewater treatment: Principles and applications: A review
.
Journal of Water Process Engineering
35
(
10
),
1
21
.
https://doi.org/10.1016/j.jwpe.2020.101193
.
Matilainen
A.
,
Vepsäläinen
M.
&
Sillanpää
M.
2010
Natural organic matter removal by coagulation during drinking water treatment: A review
.
Advances in Colloid and Interface Science
159
(
2
),
189
197
.
http://dx.doi.org/10.1016/j.cis.2010.06.007
.
Meroufel
B.
,
Benali
O.
,
Benyahia
M.
,
Benmoussa
Y.
&
Zenasni
M. A.
2013
Adsorptive removal of anionic dye from aqueous solutions by Algerian kaolin: Characteristics, isotherm, kinetic and thermodynamic studies
.
Journal of Materials and Environmental Science
4
(
3
),
482
491
.
Mohamed
M. H.
&
Wilson
L. D.
2015
Kinetic uptake studies of powdered materials in solution
.
Nanomaterials
5
(
2
),
969
980
.
https://doi.org/10.3390/nano5020969
.
Mohamed
M. H.
,
Wilson
L. D.
,
Headley
J. V.
&
Peru
K. M.
2011
Investigation of the sorption properties of β-cyclodextrin-based polyurethanes with phenolic dyes and naphthenates
.
Journal of Colloid and Interface Science
356
(
1
),
217
226
.
http://dx.doi.org/10.1016/j.jcis.2010.11.002
.
Mohammadi
E.
,
Daraei
H.
,
Ghanbari
R.
,
Dehestani Athar
S.
,
Zandsalimi
Y.
,
Ziaee
A.
,
Maleki
A.
&
Yetilmezsoy
K.
2019
Synthesis of carboxylated chitosan modified with ferromagnetic nanoparticles for adsorptive removal of fluoride, nitrate, and phosphate anions from aqueous solutions
.
Journal of Molecular Liquids
273
(
10
),
116
124
.
https://doi.org/10.1016/j.molliq.2018.10.019
.
Mohapatra
T.
,
Kumar
V.
,
Sharma
M.
&
Ghosh
P.
2021
Hybrid fenton oxidation processes with packed bed or fluidized bed reactor for the treatment of organic pollutants in wastewater: A review
.
Environmental Engineering Science
38
(
6
),
443
457
.
https://doi.org/10.1089/ees.2020.0070
.
Moreira
N. F. F.
,
Orge
C. A.
,
Ribeiro
A. R.
,
Faria
J. L.
,
Nunes
O. C.
,
Pereira
M. F. R.
&
Silva
A. M. T.
2015
Fast mineralization and detoxification of amoxicillin and diclofenac by photocatalytic ozonation and application to an urban wastewater
.
Water Research
87
(
8
),
87
96
.
https://doi.org/10.1016/j.watres.2015.08.059
.
Nejatbakhsh
S.
,
Soodmand
A. M.
,
Azimi
B.
,
Farshchi
M. E.
,
Aghdasinia
H.
&
Kazemian
H.
2023
Semi-pilot scale fluidized-bed reactor applied for the Azo dye removal from seawater by granular heterogeneous fenton catalysts
.
Chemical Engineering Research and Design
195
(
5
),
1
13
.
https://doi.org/10.1016/j.cherd.2023.05.026
.
Panigrahy
N.
,
Priyadarshini
A.
,
Sahoo
M. M.
,
Verma
A. K.
,
Daverey
A.
&
Sahoo
N. K.
2022
A comprehensive review on eco-toxicity and biodegradation of phenolics: Recent progress and future outlook
.
Environmental Technology and Innovation
27
(
2
),
1
32
.
https://doi.org/10.1016/j.eti.2022.102423
.
Qu
X.
,
Zheng
J.
&
Zhang
Y.
2007
Catalytic ozonation of phenolic wastewater with activated carbon fiber in a fluid bed reactor
.
Journal of Colloid and Interface Science
309
(
2
),
429
434
.
https://doi.org/10.1016/j.jcis.2007.01.034
.
Roshani
B.
,
McMaster
I.
,
Rezaei
E.
&
Soltan
J.
2014
Catalytic ozonation of benzotriazole over alumina supported transition metal oxide catalysts in water
.
Separation and Purification Technology
135
(
8
),
158
164
.
http://dx.doi.org/10.1016/j.seppur.2014.08.011
.
Said
K. A. M.
,
Ismail
N. Z.
,
Jama'in
R. L.
,
Alipah
N. A. M.
,
Sutan
N. M.
,
Gadung
G. G.
,
Baini
R.
&
Zauzi
N. S. A.
2018
Application of Freundlich and Temkin isotherm to study the removal of Pb(II) via adsorption on activated carbon equipped polysulfone membrane
.
International Journal of Engineering and Technology
7
(
18
),
91
93
.
https://doi.org/10.14419/ijet.v7i3.18.16683
.
Shahib
I. I.
,
Ifthikar
J.
,
Oyekunle
D. T.
,
Elkhlifi
Z.
,
Jawad
A.
,
Wang
J.
,
Lei
W.
&
Chen
Z.
2022
Influences of chemical treatment on sludge derived biochar; physicochemical properties and potential sorption mechanisms of lead (II) and methylene blue
.
Journal of Environmental Chemical Engineering
10
(
3
),
1
12
.
https://doi.org/10.1016/j.jece.2022.107725
.
Sinharoy
A.
,
Saikia
S.
&
Pakshirajan
K.
2019
Biological removal of selenite from wastewater and recovery as selenium nanoparticles using inverse fluidized bed bioreactor
.
Journal of Water Process Engineering
32
(
7
),
1
15
.
https://doi.org/10.1016/j.jwpe.2019.100988
.
Tahir
S. S.
&
Rauf
N.
2006
Removal of a cationic dye from aqueous solutions by adsorption onto bentonite clay
.
Chemosphere
63
(
11
),
1842
1848
.
https://doi.org/10.3390/w12092430
.
Tang
Q.
,
Li
N.
,
Lu
Q.
,
Wang
X.
&
Zhu
Y.
2019
Study on preparation and separation and adsorption performance of knitted tube composite β-cyclodextrin/chitosan porous membrane
.
Polymers
11
(
11
),
1
17
.
https://doi.org/10.3390/polym11111737
.
Tetteh
S.
,
Zugle
R.
,
Ofori
A.
&
Adotey
J. P. K.
2020
Kinetics and equilibrium thermodynamic studies of the adsorption of phenolphthalein and methyl orange onto muscovite clay
.
Frontiers in Chemical Research
2
(
1
),
33
37
.
https://doi.org/10.22034/fcr.2020.122175.1017
.
Tongchoo
P.
,
Intachai
S.
,
Pankam
P.
,
Suppaso
C.
&
Khaorapapong
N.
2020
The usage of CoAl-layered double oxide for removal of toxic dye from aqueous solution
.
Journal of Metals, Materials and Minerals
30
(
4
),
45
50
.
https://doi.org/10.14456/jmmm.2020.50
.
Tran
H. N.
,
You
S. J.
,
Nguyen
T. V.
&
Chao
H. P.
2017
Insight into the adsorption mechanism of cationic dye onto biosorbents derived from agricultural wastes
.
Chemical Engineering Communications
204
(
9
),
1020
1036
.
Udoetok
I. A.
,
Dimmick
R. M.
,
Wilson
L. D.
&
Headley
J. V.
2016
Adsorption properties of cross-linked cellulose-epichlorohydrin polymers in aqueous solution
.
Carbohydrate Polymers
136
(
9
),
329
340
.
https://doi.org/10.1016/j.carbpol.2015.09.032
.
Uyar
T.
,
Havelund
R.
,
Nur
Y.
,
Hacaloglu
J.
,
Besenbacher
F.
&
Kingshott
P.
2009
Molecular filters based on cyclodextrin functionalized electrospun fibers
.
Journal of Membrane Science
332
(
1–2
),
129
137
.
https://doi.org/10.1016/j.memsci.2009.01.047
.
Xiong
W.
,
Cui
W.
,
Li
R.
,
Feng
C.
,
Liu
Y.
,
Ma
N.
,
Deng
J.
,
Xing
L.
,
Gao
Y.
&
Chen
N.
2020
Mineralization of phenol by ozone combined with activated carbon: Performance and mechanism under different pH levels
.
Environmental Science and Ecotechnology
1
(
12
),
1
9
.
https://doi.org/10.1016/j.ese.2019.100005
.
Yang
Y.
,
Chen
G.
,
Murray
P.
&
Zhang
H.
2020
Porous chitosan by crosslinking with tricarboxylic acid and tuneable release
.
SN Applied Sciences
2
(
3
),
1
10
.
https://doi.org/10.1007/s42452-020-2252-z
.
Yazdi
F.
,
Anbia
M.
&
Salehi
S.
2019
Characterization of functionalized chitosan-clinoptilolite nanocomposites for nitrate removal from aqueous media
.
International Journal of Biological Macromolecules
130
(
2
),
545
555
.
https://doi.org/10.1016/j.ijbiomac.2019.02.127
.
Yu
L.
,
Bi
J.
,
Song
Y.
&
Wang
M.
2022
Isotherm, thermodynamics, and kinetics of methyl orange adsorption onto magnetic resin of chitosan microspheres
.
International Journal of Molecular Sciences
23
(
22
),
1
18
.
https://doi.org/10.3390/ijms232213839
.
Yuan
P.
,
Annabi-Bergaya
F.
,
Tao
Q.
,
Fan
M.
,
Liu
Z.
,
Zhu
J.
,
He
H.
&
Chen
T.
2008
A combined study by XRD, FTIR, TG and HRTEM on the structure of delaminated Fe-intercalated/pillared clay
.
Journal of Colloid and Interface Science
324
(
1–2
),
142
149
.
https://doi.org/10.1016/j.jcis.2008.04.076
.
Zhang
B.
,
Chen
N.
,
Feng
C.
&
Zhang
Z.
2018
Adsorption for phosphate by crosslinked/non-crosslinked-chitosan-Fe(III) complex sorbents: Characteristic and mechanism
.
Chemical Engineering Journal
353
(
7
),
361
372
.
https://doi.org/10.1016/j.cej.2018.07.092
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).

Supplementary data