Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Antimicrobial magnetic Glycyrrhiza glabra nanocomposite for decolouration of water through adsorption and photodegradation

Ankita Manchandaa, Ahmed Hussain Jawharib, Ziaul Hasanc, Nazim Hasanbd, Sneha Shuklaa, Adiba Khana, Tabrez Alam Khana and Saif Ali Chaudhry*a
aDepartment of Chemistry, Jamia Millia Islamia, New Delhi 110025, India. E-mail: schaudhry@jmi.ac.in
bDepartment of Physical Sciences, Chemistry Division, College of Science, Jazan University, P. O. Box 114, Jazan 45142, Saudi Arabia
cDepartment of Biosciences, Jamia Millia Islamia, New Delhi 110025, India
dNanotechnology Research Unit, Jazan University, P.O. Box 114, Jazan 45142, Saudi Arabia

Received 12th July 2025 , Accepted 4th October 2025

First published on 31st October 2025


Abstract

A sustainable hybrid magnetic nanocomposite based on Glycyrrhiza glabra (GG), GG/γ-Fe2O3, was synthesized via one-pot co-precipitation method, and efficiently utilized for adsorption and photocatalytic degradation of two toxic model dyes, Congo red (CR) and Nile blue (NB) dyes from water. The characterization of the GG/γ-Fe2O3 was performed by using FT-IR, P-XRD, BET-BJH, FE-SEM-EDX, TEM, SAED, XPS, TGA, UV-visible, and PL techniques. The GG/γ-Fe2O3 showed significant inhibition of bacterial and fungal growth. The inhibition statistics towards gram-(+) and gram-(−) bacteria, and fungal strains were found superior as compared to the naive plant material. The influence of adsorption parameters, on dye removal efficiency, was appraised via batch methodology. The fitting of isothermal and kinetic datasets into their respective models indicated that the adsorptive removal process was governed by the Freundlich isotherm and pseudo-second order kinetics. The Langmuir saturation capacity for CR and NB was found to be 47.50 and 15.36 mg g−1, respectively. The spontaneous and physical sorption of CR and NB was delineated to be exothermic and endothermic, respectively, from 30 to 50 °C. The band gap of the GG/γ-Fe2O3 were found 1.69 eV (indirect), and 2.30 eV (direct) which established its semiconducting design, with CR solar-degradation efficiency of 92.7%, following pseudo-first-order kinetics. The degradation intermediates and mechanism have been investigated from radical quenching experiments and high-resolution LC-MS. The GG/γ-Fe2O3 exhibited structural integrity and excellent regeneration, supported by post-treatment FT-IR analysis. The reproducibility of the optimum experimental dataset under realistic conditions, including real wastewater, co-existing ions, and dye mixtures, revealed potent application of multifunctional and cost-effective GG/γ-Fe2O3 for the efficient removal of both cationic and anionic water contaminants, as well as for reducing microbial loads.


1 Introduction

The increasing level of water pollution, owing to industrialization and agricultural expansion, is of grave environmental concern.1 Untreated or partially-treated aqueous wastewater, from various industries, is interspersed with mainstream water resources, which is detrimental to human life and aquatic organisms. Dyes are a significant class of recalcitrant toxic water contaminants emanating alongside various pigments and by-products from many industries.2 From over 11[thin space (1/6-em)]800 commercially available dyes/pigments, the annual global consumption is estimated at around 19 million metric tons. Textile (42.6%), paint and coating (27.8%), plastic (16.3%), and paper, cosmetics and ink (13.3%) industries are the major consumers of dyes.3 Around 10–15% industrial dyes are released into the environment during manufacturing and processing operations.4 Congo red, CR, a highly water-soluble anionic diazo dye is utilized in textile, paper, plastic, and printing industries.5 It is considered mutagenic, teratogenic, genotoxic, neurotoxic, cytotoxic, cutaneous, and carcinogenic.2,6 The short-term ingestion of CR can lead to skin, eye, gastrointestinal irritation, and possible blood clotting.7 Likewise, Nile blue, NB, is a cationic azine dye that finds application in the textile industry for dyeing cotton, wool, and other fabrics.8,9 It can cause sleepiness, stimulation of the digestive tract, chills, respiratory tract allergy, redness, dryness/irritation of eyes/mouth/throat/skin, dermatitis, lung cancer, and even chromosomal aberrations.8,10,11 The natural decomposition of such types of dyes is extremely difficult mainly due to their persistency, non-biodegradability, hydrophilicity, and stability.12,13 Additionally, dyes alter water transparency, impede sunlight penetration into the water bodies, fluctuate the water quality parameters, and cumulate the oxidative stress, which in turn severely affects the aquatic biota owing to decreased photosynthetic action.14 It is thereby imperative to amputate water resources of these hazardous pollutants, rendering water safe for human consumption.

Numerous techniques have been employed to clean dye-polluted water.15 However, the conventional treatment methods are cumbersome, expensive, time and energy-consuming, show low removal efficiency and residual discharge.16 Adsorption and photocatalytic degradation, of pollutants, have caught attention of many researchers, being simple, environment-friendly alternative, with minimum sludge formation, and high mineralization efficiency.17

Over last few decades, a variety of materials have been developed for the adsorptive removal of toxic dyes which include nanogels,18,19 metal-oxide nanocomposites,20–23 metal–organic frameworks,24 plant-based materials,10,25 and activated carbon/biochar26,27 because of their ease of preparation, inexpensiveness, effectiveness over a wide range of pollutants, and porous structure with high adsorption capabilities and kinetics. However, restricted functionality, regeneration, and imitation to real environmental scenarios amidst competing conditions limit their practical application. Moreover, the complete removal of some azo-dyes (e.g., CR) is rather difficult because of their complex aromatic structure, thermodynamic stability and non-biodegradability.11 The secondary by-products formed being more toxic than the parent materials, render the synthesis of novel visible-light responsive photocatalysts and porous nanocomposites inevitable.28

Therefore, adsorption coupled with heterogeneous photocatalysis has received considerable attention as an economical, rapid and reliable treatment option. Many nanostructured metal oxides, preferably iron oxides such as magnetite (Fe3O4) and maghemite (γ-Fe2O3)6,29–31 and their hybrid nanocomposites13,32 have evinced considerable interest for the removal of pernicious dyes, owing to surface defects, high surface area to volume ratio, intrinsic reactivity of surface functional sites, wide light response range, narrow band gap, chemical stability, low-cost, non-toxicity, biocompatibility, and abundancy. Further, the surface modification of nanoparticles by compositing with plant residues (biomass) substantially reduces their strong tendency to agglomerate. Compositing also prevents premature surface saturation during treatment and acts as a solid-phase co-substrate to support oxide nanoparticles.33,34 Additionally, the presence of pathogens, like bacteria and fungi, in wastewater, may cause severe health hazards, leading to various chronic and acute respiratory, gastrointestinal, and skin-related disorders.35 Therefore, in addition to synergistic dye adsorption and photodegradation, the antipathogenic activity of the synthesized composite is vital to reduce the microbial loads of the polluted water.36–38 Glycyrrhiza glabra (GG), commonly called licorice, is a traditional medicinal plant, possess therapeutic and pharmacological activities,39 and is a proficient bio-sorptive material because of its active oxygenous functional surface sites.40 It is a naturally abundant, biodegradable, and renewable bioresource, particularly, valued for its rich phytochemical composition, which bequeath antimicrobial and antioxidant potential.39 The amalgamation of GG and γ-Fe2O3 into a hybrid composite enhances removal efficiency, surface active sites, structural stability, and recovery using magnetic separation to efficaciously treat wastewater for charged contaminants, and even microbes. It is a step forward towards circular bioeconomy, repurposing waste plant/agro-biomass to enhance resource efficiency, and minimize ecological impact to synthetic chemicals, promoting sustainable environmental remediation.41 Moreover, research indicates use of iron oxide nanoparticles, particularly Fe2O3 as iron micronutrient fertilizer for improved soil texture and plant growth. It gradually increases the level of iron in soil over time.42,43 Thus, water enriched with GG and γ-Fe2O3 together can enhance soil and water dynamics after treatment.

Several synthetic strategies, including green synthesis, hydrothermal, microemulsions, sol–gel, thermal decomposition, and co-precipitation, have been developed for the production of γ-Fe2O3-based nanocomposites.33 Of these, co-precipitation is simplest, most cost-efficient, allows for better control of size and morphology, composition, stability, reproducibility, functionality, and can produce a large amount of product in a single batch.44

G. glabra (biomass and biochar) has been previously explored for water remediation,45–48 however, much attempts have not been made towards its modification into nanocomposite. Moreover, no study has schematically delved into the multifunctional approach considering adsorption, photodegradation, and antimicrobial tendency of GG composites in real wastewater conditions.

Altogether, the present study outlines co-precipitation mediated synthesis of an inexpensive advanced multifunctional magnetic hybrid nanocomposite, GG/γ-Fe2O3, which was characterized using FT-IR, P-XRD, BET-BJH, FE-SEM-EDX, TEM, SAED, XPS, TGA, UV-visible, and PL techniques. The GG/γ-Fe2O3 was explored for investigating its potential to decolourize wastewater by removing model CR and NB dyes. Different operational conditions (dosage, initial dye concentration, contact time, initial water pH and temperature) have been optimized. The sorption data was fitted into various isotherm and kinetic models, and thermodynamic equations for proposing plausible mechanism. The adsorption and degradation performance of the GG/γ-Fe2O3 towards real wastewater, and in presence of competing ions, and dye mixtures have been thoroughly investigated. The stability, and regeneration potential have also been investigated. Lastly, the antibacterial and antifungal activities of the naive GG and GG/γ-Fe2O3 against Gram-negative, E. coli (ATCC-25922), and Gram-positive, S. aureus (MTCC-902), bacterial strains, and fungal stain, C. albicans (SC-5314 and ATCC-90028), have also been investigated.

2 Materials and methods

2.1. Reagents and materials

The chemicals and instruments used for the synthesis and characterization of GG/γ-Fe2O3 have been depicted in Tables S1 and S2, respectively.

2.2. Preparation of GG/γ-Fe2O3

The GG roots were washed several times with deionized water to leach out solid or water-soluble impurities, oven-dried, and subjected to physical pre-treatment through pulverization and sieving to a variable grain diameter (mesh) size of 60–200. Such pre-treatment downsizes particle dimensions and increases the available specific surface area49 for the growth of γ-Fe2O3 nanoparticles on powdered GG following a slightly modified simple one-pot co-precipitation method.50,51 1.0 g of GG powder was dispersed in 100 mL of distilled water (DW) and ultrasonicated for 15 min to prepare a homogeneous suspension. The suspension was stirred, over a magnetic stirrer, followed by the addition of 100 mL of 0.5 M Fe(NO3)3·9H2O and 0.25 M FeSO4·7H2O each, and was subjected to continuous stirring for 30 min at 50–60 °C. This was followed by the addition of 2 M NaOH solution dropwise until the suspension turned alkaline (∼10–11 pH). The obtained brownish-black precipitate was allowed to settle, collected using a magnetic-separation method, and washed with deionized water, multiple times, followed by ethanol. It was oven-dried at 70 °C, calcined at 200 °C for 2 h, sieved, and stored in glass bottle. The schematic representation of the synthesis of GG/γ-Fe2O3 is depicted in Scheme 1.
image file: d5ra04982b-s1.tif
Scheme 1 Stepwise depiction of the synthesis of GG/γ-Fe2O3 nanocomposite.

2.3. Point of zero charge (ZPC)

The ZPC, pH at which the surface possesses zero charge,52 was determined by following a slightly modified salt addition method.53 A brief setup of a series of 100 mL Erlenmeyer flasks comprising double of the optimized dosage of GG/γ-Fe2O3 in 20 mL of 0.2 M KNO3 solutions, having pH variation in the pH range 2–10 were prepared and agitated for 24 h at 200 rpm. The final pHs of all solutions were recorded, and the point of intersection of the initial and final pHs curves established pHZPC of the GG/γ-Fe2O3.

2.4. Antimicrobial activity

The antibacterial activities of the GG and GG/γ-Fe2O3 were investigated using in vitro 96-well microtiter assays against E. coli and S. aureus bacteria. The broth micro dilution method was followed for the determination of the minimum inhibitory concentration (MIC), and minimum bactericidal concentration (MBC).54 Each test strain was cultivated overnight in nutrient broth, with the turbidity adjusted to 0.5 McFarland units in 100 μL of Mueller–Hinton broth. The GG and GG/γ-Fe2O3 stock suspensions (5.0 mg mL−1) were added to each well. Positive and negative controls were also added, and then the plates were incubated at 37 °C for 12 h. To obtain MBC values, 100 μL of turbidity-free tube content was cultured in Mueller–Hinton agar and incubated for 24 h at 37 ± 0.1 °C.55

The antifungal efficacy of the GG and GG/γ-Fe2O3, in terms of MIC and minimum fungicidal concentration (MFC)56 were determined against C. albicans (SC-5314) and C. albicans (ATCC-90028), as described in the literature.57 Briefly, 100 μL of yeast peptone dextrose media (YPD) and GG and GG/γ-Fe2O3 stock solutions (5 mg mL−1) were added to the first well, which was then serially diluted. Following the addition of 100 μL of fungal inoculant as a suspension in each well, the plates were incubated at 28 °C for 24 h. Plates without GG and GG/γ-Fe2O3 served as a negative control, while the antifungal drug, fluconazole, as a positive control. For both assays, the respective culture growth was noted as absorbance at 600 nm on Elisa plate reader. All experiments were conducted in triplicate simultaneously.

2.5. Batch adsorption study and statistical data analysis

The factors which affect the removal efficiency were optimized by varying GG/γ-Fe2O3 dosage (0.5–3 g L−1), dyes concentration (5–30 mg L−1), reaction time (15–120 min), solution pH (2–11), and temperature (30, 40, 50 °C). Briefly, in 100 mL Erlenmeyer flasks, CR and NB solutions (10 mL) having concentrations of 10 and 20 mg L−1, respectively, were mechanically shaken at 200 rpm at 30 °C and natural pH with varying dosage for 120 min. The effect of temperature was investigated using the optimized dosage of GG/γ-Fe2O3 (2.0 g L−1), and [CR] = 10 mg L−1 and [NB] = 20 mg L−1. The experiments were performed in triplicate, and the mean values have been reported.

The rationale for selecting the experimental variables and their transferability to real systems was supported by preliminary batch adsorption trials and relevant literature.53,58 The initial screening experiments, using 0.5–3 g L−1 GG/γ-Fe2O3 for 5–30 mg L−1 CR and NB, showed good adsorption results without causing saturation or particle agglomeration. The time of contact, 15–120 min, ensured sufficient sorbent–pollutant interactions and practical treatment duration, as evident from the kinetic investigation. The pH range 2–11 was chosen, considering the pH sensitivity of CR and NB dyes, to obtain the maximum adsorption. Further, the reproducibility, reliability, and statistical significance of the experimentally optimized variables were realized using one-way analysis of variance (ANOVA, Origin Pro 8.5), at the 0.05 level. The standard deviation (SD) values (Table S3) have been depicted as error in Fig. S3, while the complete statistical analysis is given in Table S4. The realized experimental variable range ensured an efficient sorptive system and statistically valid optimisation.

After adsorption and magnetic separation of the dye-loaded GG/γ-Fe2O3, the concentrations of residual dyes were determined at λmax 498 nm for CR and 627 nm for NB using a UV-visible spectrophotometer (T80+ UV/VIS, PG instruments Ltd, Leicestershire, England). The following equations were employed for the evaluation of per cent removal [eqn (1)] and equilibrium adsorption capacity [eqn (2)] of the GG/γ-Fe2O3:59

 
image file: d5ra04982b-t1.tif(1)
 
image file: d5ra04982b-t2.tif(2)
where, m (g) is the mass of GG/γ-Fe2O3, V (L) is volume of CR or NB solution; C0 and Ce are initial and final equilibrium concentrations of the dye solutions (mg L−1), respectively; and Qe (mg g−1) is the equilibrium adsorption capacity of the GG/γ-Fe2O3.

2.6. Photocatalytic degradation

Photocatalytic activity of the GG/γ-Fe2O3 against CR was investigated at its optimized sorption conditions. 2.0 g L−1 of nanocomposite was placed in 100 mL of 10 mg L−1 CR solution in dark for 1 h, under constant stirring for attainment of adsorption–desorption equilibrium on the catalytic surface. It was followed by subsequent irradiation under sunlight with uninterrupted stirring for 0–160 min. 10 mL of the degraded supernatant dye solution was withdrawn at 20 min intervals, centrifuged, and analysed spectrophotometrically. The time-dependent UV-visible spectra (A vs. λ) were recorded from λ 250 to 600 nm, for different intervals of time, and photocatalytic degradation efficiency (% DE) was evaluated using the following equation60 [eqn (3)]:
 
image file: d5ra04982b-t3.tif(3)
where, Ct (mg L−1) is the concentration of the dye solution at irradiation time t.

The kinetics of the degradation, and the rate constant of the process were computed by employing the following pseudo-first order kinetic relationship [eqn (4)]:

 
image file: d5ra04982b-t4.tif(4)
where, k is the pseudo-first order rate constant, while A0/C0 and At/Ct are the initial and final absorbance/concentration, respectively.60

Further, radical trapping experiments were performed for the detection of active degradation species.61 4.0 mL each of benzoquinone (BQ, 1 mM), isopropanol (IPA, 2[thin space (1/6-em)]:[thin space (1/6-em)]20 vol/vol), and ethylenediaminetetraacetic acid (EDTA, 10 mM) scavengers were introduced to 100 mL, 10 mg L−1 CR solution, and used to trap superoxide radicals (O2˙), hydroxyl radicals (OH˙), and holes (h+), respectively. Lastly, a plausible mechanism, for the photodegradation of CR, has been proposed in accordance with the liquid chromatography-high resolution mass spectrometry (LC-HRMS) results.

2.7. Real wastewater analysis and effect of co-existing ions

The adsorption and photodegradation tendency of GG/γ-Fe2O3 towards actual real water sample were evaluated using sewage water (SW, collected from sewage treatment plant, Batla House, New Delhi), RO water (RO), and tap water (TW) spiked with 10 mg L−1 CR and 20 mg L−1 NB dye concentrations, under optimized conditions, at their natural pH. Such water samples have ample co-existing ions which might compete with the charged organic water contaminants for sorptive sites, interfering with their removal. Thereby, 10 mM of various inorganic anions (Cl, NO3, SO42−, and CO32−), and cations (Na+, Ca2+) were selected to investigate the adsorption and degradation effect of GG/γ-Fe2O3 for CR and NB.

2.8. Adsorption selectivity in dye mixture

Real wastewater might contain a mixture of organic dyes, and the advancement of new materials for targeted adsorption relies on their selectivity. The adsorption selectivity of 2.0 g L−1 GG/γ-Fe2O3 for CR and NB dyes was justified through single-dye system, their binary mixture, and of their mixture in a quaternary setup of two anionic (A) and two cationic (C) dyes each. The concentrations of all anionic dyes were taken as 10 mg L−1, and cationic dyes as 20 mg L−1.

2.9. Regeneration study

The reusability analysis of the spent GG/γ-Fe2O3 is significant for commercial application to diminish the overall treatment cost, promoting a sustainable and efficient water treatment process. For regeneration, 2.0 g L−1 of CR/NB-loaded GG/γ-Fe2O3 nanocomposite was dispersed in 10 mL of ethanol and the suspensions were agitated in a water bath shaker at 30 °C at 200 rpm for 4 h. The spent GG/γ-Fe2O3 was washed until neutral pH, dried, and subjected to subsequent reusability–regeneration cycles. The percentage removal of dyes were determined spectrophotometrically in each cycle using eqn (1). The stability of spent GG/γ-Fe2O3 was established via FT-IR analysis post adsorption of CR and NB after a considerable number of adsorption–desorption cycles.

3 Result and discussion

3.1. Mechanism for preparation of GG/γ-Fe2O3

The mechanism of formation of GG/γ-Fe2O3 can be understood by considering the entrapment of Fe2+ and Fe3+ by oxygeneous groups (–OH and –COOH) at the GG surface through site-specific adsorption. The precipitating agent NaOH supplies enough OH which might have electrostatically attracted the adsorbed Fe2+ and Fe3+ yielding respective hydroxides, M(OH)x that ultimately transformed into multiple hydroxyls bearing Fe(OH)n2−n and Fe(OH)n3−n species.62 Afterwards, dehydration and subsequent nucleation of these hydroxides, on porous GG surface, might have led to the formation of γ-Fe2O3, and later GG/γ-Fe2O3 via incorporation with the GG framework (Fig. S1). However, it may also be stated that the high concentration of NaOH, in the reaction mixture, might have also been responsible for driving the nucleation process towards the formation of tiny nuclei, leading to a decrease in the crystallite size of particles.63 A theoretical mechanism depicting the formation of GG/γ-Fe2O3 can be proposed as [eqn (5)–(9)]:62,64,65
 
Fe2+ (aq) + Fe3+ (aq) + 5OH (aq) → Fe(OH)2 (s) + Fe(OH)3 (s) (5)
 
Fe(OH)2 (s) + (n − 2)OH (aq) → Fe(OH)n2−n (s) (6)
 
Fe(OH)3 (s) + (n − 3)OH (aq) → Fe(OH)n3−n (s) (7)
 
image file: d5ra04982b-t5.tif(8)
 
γ-Fe2O3 (s) + GG (s) → GG/γ-Fe2O3 (s) (9)

3.2. Characterization

3.2.1. Fourier transform infrared spectroscopy. The Fourier transform infrared (FT-IR) analysis of the GG root powder specified the chemical composition and bonding of various functional groups, which results in the formation of the GG/γ-Fe2O3. The FT-IR spectrum of the GG {Fig. 1A(a)} showed a large number of peaks in the range 4000–650 cm−1. A broad band, for O–H bonds due to alcoholic/phenolic compounds, like cellulose, hemicellulose or lignin, owing to intermolecular hydrogen bonding, was observed at 3362 cm−1. It was followed by strong anti-symmetric and medium symmetric C–H stretching frequencies of aliphatic –CH3 and –CH2 groups around 2930 and 2882 cm−1, respectively.66 The absorption peaks around 2136, 1637, and 1516 cm−1 corresponded to the presence of benzene rings of flavonoids, phytosterols, etc., appeared for stretching vibrations of C[triple bond, length as m-dash]C, C[double bond, length as m-dash]O in polyphenol carbonyl or carboxyl groups of flavonoids in conjugation with, or of C[double bond, length as m-dash]C of aliphatic and aromatic systems.66,67 Moreover, the peak at 1637 cm−1 also depicted the bending vibration of the O–H bond due to the entrapped moisture.68 The peaks at 1423 and 1370 cm−1 indicated the C–H bending vibrations due to –CH3 and –CH2 groups, respectively, from aliphatic chains or methoxy (O–CH3) groups in lignocellulosic materials. The bands at 1246, 1152, and 1026 cm−1 explicated deformational vibration of the C–O bond from acids, and/or ester functional groups of glycosides.66,69 While, the presence of flavonoids and saponins were confirmed by strong coupled vibrations at 849 and 762 cm−1 due to out-of-plane bending deformations of C–H bond.66 Thereby, the abundance of alcoholic, phenolic, carbonyl, carboxyl, and ester functional groups, on the GG surface, was quite evident.
image file: d5ra04982b-f1.tif
Fig. 1 (A) FT-IR spectra of (a) GG, and (b) GG/γ-Fe2O3; (B) P-XRD analysis of (a) GG, and (b) GG/γ-Fe2O3; (C) (a) BET, and (b) BJH curves; (D) MH magnetization curve of GG/γ-Fe2O3; (E) Van't Hoff plots.

The FT-IR spectrum of the GG/γ-Fe2O3 {Fig. 1A(b)} was scrutinized to discern the nature of interactions between GG surface functional groups and the integrated oxides. Slightly shifted and less intense peaks, with usual annotations to GG, were ascertained at 3378, 2925, 2854, 1624, 1462, 1381, 1022, 887, and 794 cm−1. The peak at 1152 cm−1 showed no shifting; however, some of the minor peaks got quenched, and some new peaks were spotted. This behaviour could be attributed to the possible increase in bond strength and formation of electrostatic and hydrogen bonds. Some new peaks, in the range 650–400 cm−1, depicted the successful formation of GG/γ-Fe2O3 via metal–oxygen bond vibrations. Peaks at 623, 589, and 449 cm−1 might be attributed to the Fe–O–C stretching vibrations which are characteristics of ν(Fe–O) stretching mode in iron oxides.68,70

3.2.2. Powder-X-ray diffraction crystallography. The Powder-X-ray diffraction (P-XRD) patterns of GG and GG/γ-Fe2O3 (Fig. 1B) were analyzed in the angular range 5–90° (2θ). A broad peak around 22.48° (2θ) was observed for the GG corresponding to the (002) plane, characteristic peak of amorphous cellulosic material.71 However, the diffraction pattern of the GG/γ-Fe2O3 manifested distinct intense peaks at 18.14° (111), 30.1° (220), 35.54° (311), 43.46° (400), 53.98° (422), 57.16° (511), 62.76° (440), 70.9° (620), 73.98° (533), 78.86° (622) 21.14° (200), 23.78° (210), 26.10° (211), 32.24° (300), and 39.36° (320) due to the presence of γ-Fe2O3 (JCPDS PDF-04-0755). Moreover, a gradual change in colour from black to brownish black on heating exemplified the slow oxidation of magnetite to maghemite.72 The angular values ruled out the possibility of formation of other types of iron oxides, i.e., hematite (α-Fe2O3), goethite [FeO(OH)], wüstite (FeO), etc.73 Meanwhile, the occurrence of most intense peak at 35.54°, corresponding to (311) plane, was close to the standard angular value of 35.597° for γ-Fe2O3, signifying γ-Fe2O3 as the predominant crystalline phase in GG/γ-Fe2O3.22 On comparison of the two diffractograms (Fig. 1B) it was observed that GG/γ-Fe2O3 displayed peaks of both GG and γ-Fe2O3 at various 2θ levels. However, the intensity of diffraction peak at 22.48° for GG significantly got reduced in GG/γ-Fe2O3 owing to chemical interactions between its oxygeneous functional groups and γ-Fe2O3 nanoparticles. The mean crystallite size, corresponding to prominent peaks of the XRD pattern, was estimated by fitting the angular data in the Debye–Scherrer equation (eqn (S1))74 as 36.09 nm, confirming the nanostructure of high surface area, suitable for greater dye adsorption. Moreover, the nature and intensity of the peaks indicated low crystallinity of GG/γ-Fe2O3.
3.2.3. BET and BJH analysis. Brunauer–Emmett–Teller (BET) and Barrett–Joyner–Halenda (BJH) analysis was followed to determine surface area, average pore width, and single point total pore volume (of pores less than 413.690 Å width at P/P0 = 0.951) of the GG/γ-Fe2O3, which were found to be 114.454 m2 g−1, 71.549 Å, and 0.205 cm3 g−1, respectively. The amount of the N2 gas adsorbed by the GG/γ-Fe2O3 decreased with reducing pressure, and N2 adsorption–desorption isotherm was similar to Type-IV, characteristic of a mesoporous structure {Fig. 1C(a)}.75 The adsorption average pore size (4 V A−1) distribution was centered at 75.251 Å, while the desorption average pore size (4 V A−1) was found to be 70.258 Å from the BJH plot {Fig. 1C(b)}. The specific surface area and porosity of the GG/γ-Fe2O3 were found to be appreciably high, indicating superior physical and textural characteristics.
3.2.4. Vibrating sample magnetometry. The magnetic hysteresis, MH curve, recorded via vibrating sample magnetometry (VSM), indicated the magnetic nature of the GG/γ-Fe2O3 in field strength of −2 T to +2 T (Fig. 1D). An S-shaped curve was observed, which suggested superparamagnetism, at room temperature mainly of the single magnetic domains of the nanoparticles in the material.76 The saturation magnetic moment, Ms, of the GG/γ-Fe2O3 was found to be 26.827 emu g−1, which is modest for the easy magnetic separation of the nanocomposite from the treated water.
3.2.5. FE-SEM-EDX and TEM. Field emission scanning electron microscopy (FE-SEM) imaging revealed a smooth and even surface of GG (Fig. 2a) and the introduction of heterogeneity and roughness in the GG/γ-Fe2O3 surface due to clustering of irregular oxide nanoparticles (Fig. 2b and c). An enhancement in overall contending porosity and surface area was observed in GG/γ-Fe2O3 for increased CR and NB uptake. The CR accumulation on the GG/γ-Fe2O3 surface (Fig. 2d) caused abatement of even surface with significant accretion making the surface inhomogeneous. Elemental composition of the GG and GG/γ-Fe2O3 were ascertained from the energy dispersive X-ray spectroscopy, EDX, and elemental mapping (Fig. 3a and b). The normalized atomic percentage of 67.70% C, and 32.30% O were detected for GG, while the composition fluctuated to 23% C, 64% O, and 13% Fe owing to the nanostructured growth of γ-Fe2O3 on GG framework. The analysis suggested the presence of C, H and O due to GG, while increase in % O, decrease in % C, and presence of iron was due to formation of nanostructured Fe(III) oxide which deposited in the form γ-Fe2O3.
image file: d5ra04982b-f2.tif
Fig. 2 FE-SEM images of (a) GG, (b) GG/γ-Fe2O3, (c) GG/γ-Fe2O3 before adsorption, and (d) GG/γ-Fe2O3@CR.

image file: d5ra04982b-f3.tif
Fig. 3 EDX and mapping of (a) GG and (b) GG/γ-Fe2O3.

Transmission electron microscopy (TEM) (Fig. 4a and c) provided a clear indication of the average size distribution of the GG/γ-Fe2O3 particles. The mean particle size of the GG/γ-Fe2O3 (Fig. 4b) was estimated around 31.92 nm, which is in line with the XRD result (Fig. 1B). The contrast between the brighter GG framework and darker patches of γ-Fe2O3 corresponded to the uniform deposition, entrapment, and random dispersion of nanoparticles throughout. Moreover, the TEM images confirmed the amorphous and heterogeneous nature of GG/γ-Fe2O3. The lattice fringes with protuberant electron diffraction rings in the selected area electron diffraction (SAED) pattern (Fig. 4d) indicated an amorphous nature with slight crystallinity of γ-Fe2O3 nanoparticles in GG/γ-Fe2O3.77 The d-spacing values and corresponding Miller indices of the six diffraction rings (Table S5) aligned with the XRD findings.


image file: d5ra04982b-f4.tif
Fig. 4 (a)–(c) represent TEM images, and (d) the SAED ring pattern of GG/γ-Fe2O3.
3.2.6. X-ray photoelectron spectroscopy. The chemical environment and the elemental valence states were further investigated through X-ray photoelectron spectroscopy (XPS) analysis of the GG/γ-Fe2O3 (Fig. 5). The co-existence of elements Fe, O, and C was confirmed, which was consistent with the EDX data. The high-resolution narrow spectrum of Fe 2p (Fig. 5b) matches the core-level binding energies of the spin–orbit doublets Fe 2p3/2 and Fe 2p1/2, centered at 710.4 and 724 eV, which is characteristic of Fe3+ in γ-Fe2O3. The Fe3+ 2p3/2 can be deconvoluted into two distinct sub peaks at 710.2 and 712 eV for Fe3+ in octahedral and tetrahedral sites, respectively.78 Further, shake-up satellite peaks for Fe 2p3/2 at 718.9, and Fe 2p1/2 at 73379 were indicative of transition of Fe-3d electrons to empty 4s orbital during ejection of core 2p photoelectrons.78 Moreover, the O 1s scan of GG/γ-Fe2O3 (Fig. 5c) validated the prevalence of Fe–O bonds via peak at binding energy 529.6 eV; and the peak at 531 eV was assigned to C–O–Fe bond linkage between GG carbon framework and γ-Fe2O3.80 An additional peak at 532 eV was attributed to the C–O units (C–OH/C–O–C) in oxygen-containing functional groups.81 Lastly, the C 1s XPS spectrum (Fig. 5d) exhibited three fitted peaks of C–C/C[double bond, length as m-dash]C, C–O, and O–C[double bond, length as m-dash]O centered at 284.6, 286.18 and 288.2 eV.82 These results validated the FT-IR established functional groups and the formation of new metal–oxygen bonding interactions in the GG/γ-Fe2O3 nanocomposite.
image file: d5ra04982b-f5.tif
Fig. 5 (a) The full survey scan, and high-resolution spectra of (b) Fe 2p, (c) O 1s, (d) C 1s of GG/γ-Fe2O3; (e) TGA of GG/γ-Fe2O3; (f) PL spectra of γ-Fe2O3 and GG/γ-Fe2O3.
3.2.7. Thermogravimetric analysis (TGA). The weight loss curve of the GG/γ-Fe2O3, due to thermal decomposition (Fig. 5e), shows gradual weight loss of 6.13% from room temperature to around 190 °C, related to the removal of residual surface water molecules. The primary decomposition of the sample occurred around 200–470 °C, accompanying biomass pyrolysis.83 A slight weight change was observed at higher temperature, more than 470 °C due to phase transition.84 The residual weight percent of 77.63% was realized, indicating high structural and thermal stability of the nanocomposite.
3.2.8. Optical properties of GG/γ-Fe2O3. The UV-visible spectroscopic analysis of the GG/γ-Fe2O3 (Fig. S2a) showed peaks between 320–420 nm, which are typically associated with the surface plasmon resonance of iron in γ-Fe2O3.85 The band gap energy, Eg, is crucial for predicting the photochemical and photophysical characteristics of material that is used for photocatalytic dye degradation. γ-Fe2O3 based semiconductors have been known to exhibit both indirect [O2−(2p) → Fe3+(3d)] and direct [Fe3+(3d → 3d)] transitions.86–88 The indirect and direct transition band gap of the GG/γ-Fe2O3, were determined by the Tauc plot, and found 1.69 and 2.30 eV, respectively (Fig. S2b and c). These band gap energy values established its semiconducting nature.

The photoluminescence (PL) spectra of the GG (Fig. S2d), GG/γ-Fe2O3 and pure γ-Fe2O3 (Fig. 5f) were acquired at an excitation wavelength of 325 nm to comprehend the charge carrier separation efficiency and the recombination rate of electron–hole pair. The emission varied from 350 to 800 nm, with broad and intense emission bands centered at ∼480 nm and ∼565 nm in pure γ-Fe2O3 nanoparticles which might be due to exciton emission, and radiative recombination of mobile and trapped electrons on octahedral and tetrahedral sites of γ-Fe2O3, respectively.89 However, the intensity of the peaks reduced significantly in GG/γ-Fe2O3, confirming the hampering of recombination of photogenerated charge carriers. The PL spectra of GG was of highest intensity, which reduced drastically on combination with γ-Fe2O3.

3.3. Inhibition of growth of microorganisms

The GG and GG/γ-Fe2O3 exhibited dose-dependent inhibition of pathogenic bacteria. The MICs of GG were found to be 550.25 ± 1.85 and 425.5 ± 1.25 μg mL−1, whereas for GG/γ-Fe2O3, these values were found to be 250 ± 3.15 and 200 ± 2.62 μg mL−1, for E. coli and S. aureus, respectively (Table S6). The lower MIC values of the GG/γ-Fe2O3 against both Gram-negative and Gram-positive bacteria, in comparison to GG, suggested the superiority of GG/γ-Fe2O3 for treating bacterial pathogens in wastewater. The GG and GG/γ-Fe2O3 generally showed more antibacterial effects against Gram-positive bacteria, due to differences in cell composition and thickness, compared to Gram-negative bacteria. The findings were aligned with previous studies that also indicated the antimicrobial potential of the GG.90,91

In addition, the GG and GG/γ-Fe2O3 also exhibited significant antifungal activity against two Candida albicans strains. The GG exhibited MICs of 300.85 ± 8.58 μg mL−1 and 320.74 ± 5.65 μg mL−1, while GG/γ-Fe2O3 showed MICs of 131.25 ± 2.56 μg mL−1 and 125 ± 1.15 μg mL−1, for C. albicans (SC5314) and C. albicans (ATCC90028), respectively, and separately (Table S7). The lower MIC and MFC values of the GG/γ-Fe2O3 indicated a more efficient and potent antifungal nature. Thereby, it could be justified that the enhanced antibacterial and antifungal activity of the GG/γ-Fe2O3 could be attributed to the combined effect of phytogenic contents in GG and iron oxide nanoparticles, which might had diffused and interacted with the bacterial and fungal lipid layer within their cell membrane.92

3.4. CR and NB sorption investigation

3.4.1. Influence of operational parameters on the dye removal efficiency of GG/γ-Fe2O3. At first, with increase of GG/γ-Fe2O3 dose, from 0.5 to 2.0 g L−1, a relative increase in adsorption of CR, from 68.58 to 90.49%, and NB, from 68.92 to 95.90%, was observed (Fig. S3a). The observation could be justified due to the enhancement of the overall specific surface area of GG/γ-Fe2O3, which is highly porous in nature. The FT-IR spectrum confirmed the presence of oxygenous adsorptive sites on the GG/γ-Fe2O3 surface which increased up to 2.0 g L−1 for both dyes. Following, at higher doses, the reduction in the amount of CR and NB against largely available adsorption active sites saturated the surface.22,93

2.0 g L−1 of the GG/γ-Fe2O3 efficaciously removed around 97% CR from 10 mg L−1 solution and 79% NB from the 20 mg L−1 solution. At higher concentrations, the percentage adsorption reduced slightly to 96% and 72% for 30 mg L−1 CR and NB, respectively (Fig. S3b). This observation could be attributed to the saturation of the available active sites on the otherwise fixed amount of GG/γ-Fe2O3, which prevented further adsorption. Thus, as low as 10 mg L−1 CR and 20 mg L−1 NB concentrations were optimized, since the released dye concentration in industrial effluent is not too high.

pH of water influence electrostatic interactions, directs the surface charge of solid, and the extent of ionization of dyes in water.94 Numerous surface functional sites (–COOH, –OH) are subjected to modification, i.e., either through protonation or deprotonation at pH below and above ZPC, bestowing either positive (–COOH2+, –OH2+) or negative (–COO, –O) charge to the surface, respectively, as depicted below [eqn (10)–(13)]:

At acidic pH

 
GG/γ-Fe2O3–OH (surface) + H+ (aq) ↔ GG/γ-Fe2O3–OH2+ (surface) (10)
 
GG/γ-Fe2O3–COOH (surface) + H+ (aq) ↔ GG/γ-Fe2O3–COOH2+ (surface) (11)

At alkaline pH

 
GG/γ-Fe2O3–OH (surface) + OH (aq)↔ GG/γ-Fe2O3–O (surface)+ H2O (l) (12)
 
GG/γ-Fe2O3–COOH (surface) + OH (aq) ↔ GG/γ-Fe2O3–COO (surface) + H2O (l) (13)

The GG/γ-Fe2O3 showed higher uptake capacity for CR in acidic medium, however, an insignificant gradual decrease in percentage sorption was observed on transition to neutral conditions, which further steepened in alkaline environment, i.e., 59.07% at pH = 11.95 Moreover, NB removal was at its lowest (40.36%) in highly acidic conditions (pH = 2) and increased significantly from pH 4 to 8. The maximum adsorption of NB onto GG/γ-Fe2O3 was obtained at pH 11 (Fig. S3c).

The sorption trend of GG/γ-Fe2O3 can be explained on the basis of the ZPC of 6.9, indicating a nearly amphoteric surface (Fig. S4). In acidic pH (pH < ZPC), the GG/γ-Fe2O3 exhibited greater CR adsorption owing to extensive electrostatic interactions between the protonated positively charged GG/γ-Fe2O3 surface and negatively charged CR ions. Additionally, around pH 8, hydrogen bonding between surface functional groups of the GG/γ-Fe2O3 and –NH2 group in CR was majorly governing its uptake.22 Conversely, at higher pH, the coulombic repulsion between the negatively charged deprotonated surface and the anionic CR relinquished its uptake.

Reversibly, the cationic NB exhibited favourable sorption at higher pH (pH > ZPC) following deprotonation of the GG/γ-Fe2O3 surface by OH, bestowing negative charge to the surface. Moreover, the extensive protonation of the GG/γ-Fe2O3 at low pH (pH < ZPC) decreased the sorption of positively charged NB. Further, there might be a sense of competition between H+/H3O+ and NB ions for binding sites at low pH, resulting in slower NB sorption. Therefore, a combination of electrostatic and hydrogen bonding interactions might be responsible for the adsorptive removal of CR and NB by GG/γ-Fe2O3.

Monitoring of contact time, for 10 mg L−1 CR and 20 mg L−1 NB sorption onto 2.0 g L−1 GG/γ-Fe2O3, elucidated a continuous increment in the percentage adsorption with increasing time (Fig. S3d). The adsorption of CR and NB by GG/γ-Fe2O3 proceeded in two phases: an instantaneous initial fast phase, followed by a later slow phase. The rate of removal of CR and NB showed a steady increase till 60 min (optimum time), removing nearly 88% CR and 93% NB within the first 15 min, increasing gradually to 94% CR and 96% NB removal at 60 min, after which the transport rate diminished, and nearly became constant for a prolonged time of contact. The trend can be explained based on of the availability of abundant vacant active sites (with COO and OH groups) initially, providing an easy pathway for interaction between dye molecules and 2.0 g L−1 GG/γ-Fe2O3. With time, the need for specific pathways subjugated the adsorption rate onto the partially available or later filled sites, attaining equilibrium.96

3.4.2. Effect of temperature and thermodynamics. The CR removal by the GG/γ-Fe2O3 decreased with increasing temperature, which can be attributed to an increase in mobility and a decrease in diffusion rates of CR into pores.97 Furthermore, with an increase in temperature, the weakly-bonded CR ions might have escaped from their binding sites in response to bond breakage. For instance, weak hydrogen bonds are susceptible to breakage upon receiving energy at 40 and 50 °C, resulting in reduced adsorption capacity at higher temperatures.98 Thereby, CR adsorption onto GG/γ-Fe2O3 surface was exothermic in nature.99 On the other hand, NB adsorption showed an increase with the rise in temperature, indicating an endothermic process.100 The activation of GG/γ-Fe2O3 surface at high temperatures made the adsorptive sites readily available for NB sorption.97

Adsorption thermodynamics was studied by appraising the changes in enthalpy (ΔH), entropy (ΔS), and Gibbs free energy (ΔG). Substituting the experimental data in thermodynamic equations (eqn (S2) and (S3))101,102 delivered negative values of ΔG, i.e., −6.928, −5.076, and −4.552 kJ mol−1 for 10 mg L−1 CR, and −1.579, −2.151, and −4.501 kJ mol−1 for 20 mg L−1 NB sorption, at 30, 40, and 50 °C, respectively (Table 1). These values indicated the thermodynamic feasibility of CR and NB sorption onto GG/γ-Fe2O3 within the tested temperature range. Besides, ΔG, between 0 and −20 kJ mol−1, suggested physisorption of CR and NB dyes onto the surface of GG/γ-Fe2O3.103 The surface charge significantly influences the thermodynamics of physical adsorption by altering the interaction energy, electrostatic interactions around ZPC, and strength of the weak bonding (van der Waals) interactions between the dye molecules and the interface, posing substantial impact on the overall extent of adsorption.104,105 The balance between physisorption and electrostatic interactions drives the whole sorption process. The electrostatic attractions enhance physisorption and make ΔG more negative, while electrostatic repulsions decrease physisorption and increase the ΔG value. On increasing temperature, the order of ΔG for CR removal became less negative, while a corresponding increase in negative value was observed for NB removal, suggesting a decrease in the adsorption rate in the former and a corresponding increase in the latter. For NB adsorption, a higher negative ΔG at elevated temperatures indicated greater spontaneity and affinity due to electrostatic interactions between NB and GG/γ-Fe2O3 at 50 °C. This indicated endothermic adsorption of NB which could be justified from the positive value of ΔH (+42.408 kJ mol−1). Moreover, a positive ΔS (+0.144 kJ mol−1 K−1) value suggested good affinity of NB towards GG/γ-Fe2O3, and increased randomness at the dye–adsorbent interface (Table 1). Van't Hoff plot (Fig. 1E) produced negative ΔH, −43.146 kJ mol−1, confirming exothermic thermodynamics for CR adsorption. Moreover, for CR sorption, the plot also showed a decrease in randomness, ΔS = −0.120 kJ mol−1 K−1, at the interface, indicating adsorption accompanied by a corresponding reduction in the degrees of freedom at the solid–liquid interface. It might be due to a probable increase in affinity between CR and GG/γ-Fe2O3 surface through van der Waals, electrostatic and hydrogen bonding interactions. However, the magnitude of ΔH, less than 80 kJ mol−1 for both CR and NB adsorption, supported weak physical interaction with GG/γ-Fe2O3 surface.106

Table 1 Non-linear isotherm and thermodynamic parameters for CR and NB adsorption onto GG/γ-Fe2O3
  Parameter CR NB
30 °C 40 °C 50 °C 30 °C 40 °C 50 °C
Langmuir isotherm image file: d5ra04982b-t8.tif Q0 (mg g−1) 39.744 39.789 47.504 15.361 14.902 13.305
b (L mg−1) 0.412 0.188 0.118 0.259 0.365 1.508
RL 0.195 0.347 0.459 0.162 0.120 0.032
R2 0.986 0.994 0.993 0.996 0.980 0.990
χ2 0.284 0.113 0.112 0.036 0.221 0.142
Freundlich isotherm image file: d5ra04982b-t9.tif kF [(L mg−1)1/n (mg g−1)] 11.455 6.261 5.092 3.544 4.313 7.063
1/n 0.756 0.778 0.815 0.529 0.481 0.351
n 1.323 1.286 1.227 1.892 2.077 2.845
R2 0.996 0.998 0.998 0.990 0.997 0.953
χ2 0.082 0.029 0.039 0.099 0.037 0.653
Temkin isotherm Qe = βT[thin space (1/6-em)]ln(kTCe) kT (L mg−1) 11.712 4.559 3.453 3.002 5.633 17.264
bT (kJ mol−1) 0.547 0.532 0.541 0.793 0.928 0.993
R2 0.891 0.925 0.921 0.981 0.950 0.997
χ2 2.149 1.336 1.359 0.196 0.554 0.038
D–R isotherm image file: d5ra04982b-t10.tif kD–R 0.628 1.452 2.019 3.004 2.552 0.474
QD–R (mg g−1) 15.894 14.158 13.849 9.853 10.506 11.185
ED–R (kJ mol−1) 2.248 1.527 1.336 1.028 1.152 2.758
R2 0.902 0.892 0.881 0.843 0.809 0.928
χ2 1.931 1.926 2.031 1.607 2.130 0.993
ΔG (kJ mol−1)   −6.928 −5.076 −4.552 −1.579 −2.151 −4.501
ΔH (kJ mol−1)     −43.146     +42.408  
ΔS (kJ mol−1 K−1)     −0.120     +0.144  


3.4.3. Modelling of adsorption isotherms. The sorption data, at 30, 40, and 50 °C for CR and NB (5–30 mg L−1), and 2.0 g L−1 GG/γ-Fe2O3, were fitted into different isotherm models to decipher the corresponding isotherm parameters, which delineate the surface characteristics, interactions, and sorption capacity (Table 1). The non-linear regression analysis was conducted using OriginPro 8.5 software. The statistical treatment of experimental data concerning error analysis and conformity to various models was performed using correlation coefficient (R2) and reduced chi-square (χ2), represented by eqn (14) and (15).
 
image file: d5ra04982b-t6.tif(14)
 
image file: d5ra04982b-t7.tif(15)
where Qe(cal) and Qe(meas) are the calculated (theoretical) and measured (experimental) adsorption capacities (mg g−1), respectively, [Q with combining macron]e(meas) (mg g−1) is the average of Qe(meas), p represents the number of model parameters, and N is the number of experimental data points.

3.4.3.1. Langmuir isotherm. The Langmuir isotherm assumes homogeneous, one molecule per site, monolayer adsorption onto finite and degenerate adsorptive sites via physical or chemical forces, without lateral interactions.107 Non-linear Langmuir isotherm108 expounded maximum adsorption capacities, Q0, in the range 39.744–47.504 mg g−1 for CR (increasing) and 15.361–13.305 mg g−1 for NB (decreasing) sorption, representing endothermic and exothermic adsorption, respectively, within 30–50 °C. This observation deviated from the thermodynamic interpretations. Moreover, the corresponding values of Langmuir constant, b, representing the extent of affinity in terms of binding energy was determined from the non-linear Langmuir isotherm plots (Fig. S5a and b), which decreased from 0.412 to 0.118 L mg−1 for CR; increased from 0.259 to 1.508 L mg−1 for NB removal, suggesting lower heat of adsorption/affinity of CR, and higher heat of sorption/affinity of NB with increasing temperature, which was also not in accordance with ΔS values. These interpretations suggested disagreement between the experimental data and the derived Langmuir parameters (Table 1).

In addition, the separation factor, image file: d5ra04982b-t11.tif specifies the shape of an isotherm and the reversibility of a process. If 0 < RL < 1, then the sorption is considered feasible and energetically favourable, whereas RL = 0 implies an irreversible adsorption phenomenon, and RL = 1 indicates linearity.109 RL values between 0 and 1 confirmed the energetically feasible adsorption of CR and NB onto the GG/γ-Fe2O3 surface within the tested temperature range.


3.4.3.2. Freundlich isotherm. The Freundlich isotherm assumes surface as heterogeneous and physical adsorption onto energetically non-uniform sites with explicit bond energies, resulting in multilayers via lateral interactions. It considers an exponential decrease in sorption energy upon surface coverage.110 Non-linear, Qe vs. Ce, Freundlich plots, for CR (Fig. 6a) and NB (Fig. 6b) adsorption onto GG/γ-Fe2O3, demonstrated best-fit statistics to the experimental sorption data, in terms of R2 nearest to unity. This extraction could be further confirmed from the texture of GG/γ-Fe2O3 particles in the SEM micrographs (Fig. 2b) which indicated a heterogeneous multi-layer formation. The Freundlich parameter, kF, decreased from 11.455 to 5.092 (L mg−1)1/n (mg g−1) for CR, and increased from 3.544 to 7.063 (L mg−1)1/n (mg g−1) for NB adsorption in the temperature range 30–50 °C, indicating low adsorption capacity for CR (exothermic), and high capacity for NB (endothermic) sorption at elevated temperatures. This conclusion was found in accordance with the thermodynamic data set, dictating the Freundlich isotherm fitting for both CR and NB dye removal data. The heterogeneity factor, n, obtained in the range 1–10 (n > 1) and 1 < 1/n < 0 (Table 1), established favourable physical adsorption with strong interaction of CR/NB and the GG/γ-Fe2O3 surface.110
image file: d5ra04982b-f6.tif
Fig. 6 Non-linear Freundlich isotherm plots for (a) CR, and (b) NB adsorption; non-linear pseudo-second order plots for (c) CR, and (d) NB adsorption.

3.4.3.3. Temkin isotherm. The Temkin isotherm considers heterogeneous solute–solid interactions. It assumes orderly distribution of binding energy to a certain extent, and linear decrease in adsorption enthalpy with surface coverage.111 Non-linear plots of Temkin isotherm, for CR (Fig. S5c) and NB (Fig. S5d) adsorption onto the GG/γ-Fe2O3, produced parameters kT (L mg−1) and bT (kJ mol−1), where bT (=RT/βT) is associated with heat of adsorption, and βT is the Temkin constant. The kT corresponds to the maximum binding energy and is the binding constant. The respective decrease (from 11.712 to 3.453 L mg−1) and increase (from 3.002 to 17.264 L mg−1) in parameter kT, from 30–50 °C, for CR and NB adsorption reinforced the exothermic and endothermic nature of the sorption processes, respectively. The considerable difference in kT between 30 and 40 °C indicated significant variation in the binding capabilities for CR adsorption. Moreover, the closeness in bT values for CR directed a similar extent of binding probability and constant sorption enthalpy at all temperatures. Conversely, for NB sorption, a notable increase in kT was observed between 40 and 50 °C, indicating a change in bonding pattern. It can be stated that the reported typical range of bonding energy for ion exchange mechanism is between 8 and 16 kJ mol−1,112 however, the observed low values (bT < 8 kJ mol−1) indicated weak van der Waals interactions between CR/NB and GG/γ-Fe2O3.
3.4.3.4. Dubinin–Radushkevich isotherm. The Dubinin–Radushkevich (D–R) isotherm illustrates pore-filling nature of sorption for intermediary to highly concentrated systems, and supports mechanism that follows heterogeneous Gaussian energy distribution onto specific adsorptive sites.113 The activity coefficients, kD–R (=βR2T2), for both CR and NB sorption at 30, 40, and 50 °C were interpreted, from non-linear Qe vs. Ce D–R plots (Fig. S5e and f). A decrease in the maximum (theoretical) equilibrium monolayer sorption capacity, QD–R, (from 15.894 to 13.849 mg g−1), and free energy, ED–R (from 2.248 to 1.336 kJ mol−1) with temperature was observed for CR sorption that indicated an exothermic nature, where ED–R = (2β)−0.5, and β (kJ−2 mol2) is a constant related to adsorption energy. However, both QD–R and ED–R values slightly increased in the range 9.853–11.185 mg g−1 and 1.028–2.758 kJ mol−1, respectively, with a rise in temperature from 30–50 °C, indicating endothermic NB adsorption onto the GG/γ-Fe2O3. The D–R isotherm also gave low values of ED–R, < 8 kJ mol−1, suggesting physisorption of CR and NB onto the GG/γ-Fe2O3 at three test temperatures.

Furthermore, the comparison of χ2, for various non-linear isotherms, established the lowest value for the Freundlich isotherm for both CR and NB sorption, in addition to the highest R2 values for CR adsorption, and high values for NB adsorption as well. The fitting of experimental sorption data was more pronounced for the Freundlich isotherm compared to the Langmuir, Temkin, and D–R isotherms. Thereby, adsorption of both CR and NB from their aqueous solution onto GG/γ-Fe2O3 was satisfactorily described by the Freundlich isotherm, reflecting physical nature of sorption on the energetically heterogeneous surface of GG/γ-Fe2O3. Appearance of heterogeneity might be a result of non-homogeneous distribution of adsorptive sites on the GG/γ-Fe2O3 surface, which was cross-referred and found in order with the SEM (Fig. 2b), and TEM micrographs (Fig. 4).

3.4.4. Investigation of sorption kinetics. An insight into the kinetics of the sorption reaction was devised from the dynamic parameters that regulate the overall rate of transportation of CR/NB ions from the aqueous phase onto the surface of GG/γ-Fe2O3. The kinetic data, for 15–120 min duration at 30 °C for 10 mg L−1 CR and 20 mg L−1 NB solutions with 2.0 g L−1 GG/γ-Fe2O3, were monitored at intervals of 15 min, and then fitted into pseudo-first order (PFO), pseudo-second order (PSO), Elovich (Table 2), intraparticle diffusion (IPD), and Boyd's liquid film diffusion (LFD) models (Table S8). The adequacy of each model was examined from R2 and χ2 values [eqn (14) and (15)].
Table 2 Kinetic parameters derived from non-linear kinetic plots for CR and NB adsorption onto GG/γ-Fe2O3
Pollutant Pseudo-first order Qt = Qe(1 − ek1t) Pseudo-second order

image file: d5ra04982b-t12.tif

Elovich

image file: d5ra04982b-t13.tif

Qe (cal.) (mg g−1) k1 (min−1) R2 χ2 Qe (exp.) (mg g−1) Qe (cal.) (mg g−1) k2 (g mg−1 min−1) R2 χ2 α (mg g−1 min−1) β (g mg−1) R2 χ2
CR 4.701 0.183 0.484 0.009 4.701 4.817 0.137 0.876 0.002 1.995 × 108 5.324 0.960 7.227 × 10−4
NB 9.598 0.241 0.435 0.009 9.614 9.702 0.161 0.854 0.002 3.867 × 1021 5.802 0.952 7.311 × 10−4



3.4.4.1. Pseudo-first order kinetic model. Lagergren and Ho's pseudo-first order kinetic model (PFO) assumes that the number of available or unoccupied adsorption sites is the sole governing parameter of the rate of reaction on the solid surface in a liquid–solid system. The parameters, k1 (min−1), PFO rate constant; Qe and Qt (mg g−1), the equilibrium sorption capacity, and sorption uptake at time t, respectively, were determined from the non-linear plot of Qt and t.114 The plots (Fig. S6a and b) produced comparable calculated and experimental Qe values, but with inferior R2. The rate constant k1 was found to be 0.183 and 0.241 min−1 for CR and NB removal, respectively (Table 2).
3.4.4.2. Pseudo-second order kinetic model. Ho and McKay's pseudo-second order kinetic model (PSO) assumes that in addition to the adsorption surface sites, the concentration of pollutant in the aqueous phase also determine the rate of the whole sorption process. At equilibrium, the rate-determining step (RDS) exemplifies the square of the difference between the total sorption sites and the unoccupied ones. The nature of chemical interaction between the pollutant and the surface sites controls the process.103 Qt vs. t plots (Fig. 6c and d) gave the pseudo-second order rate constant, k2 (g mg−1 min−1), and equilibrium sorption capacity, Qe.114 The parameters obtained from these plots showed fair agreement between the experimental Qe (4.701 mg g−1 for CR, and 9.614 mg g−1 for NB) and calculated Qe (4.817 mg g−1 for CR, and 9.702 mg g−1 for NB) values, greater R2, and lowest χ2 values in comparison to PFO plots (Table 2). Therefore, there was a clear-cut indication in favour of site-specific chemical interactions between CR or NB and functional groups on the GG/γ-Fe2O3 surface following PSO kinetics. This observation could be justified from the accuracy of the established resemblance between the PSO kinetic equation and the universal rate law for a chemical reaction,115 alongwith the FT-IR interpretations (Fig. 1A).
3.4.4.3. Elovich kinetic model. The Elovich model focuses on the chemical interaction between the functional sites on the liquid–solid interface. The model assumes energetically heterogeneous adsorption without sideways interaction.116 The Elovich coefficients, α (mg g−1 min−1) and β (g mg−1), exemplify the initial rate of adsorption and desorption, respectively. Moreover, β also signifies the extent of surface coverage and corresponding energy of activation for the chemisorption process. The non-linear Qt vs. t plots,117 for Elovich model (Fig. S6c and d), provided value of α parameter (1.995 × 108 mg g−1 min−1 for CR, and 3.867 × 1021 mg g−1 min−1 for NB) much greater than β (5.324 g mg−1 for CR, and 5.802 g mg−1 for NB) (Table 2), suggesting viability and feasibility of CR and NB sorption at the GG/γ-Fe2O3 specific sites with higher rate of adsorption than desorption.
3.4.5. Adsorption mechanism. Various factors influence the adsorption mechanism, including the nature of the solid surface, active sites, functionality, charge, and structure of the dyes, as well as interactions between the solid and pollutant interfaces.103 The shifting, weakening, and appearance of distinctive peaks in the FT-IR spectra of dye-loaded GG/γ-Fe2O3, i.e., GG/γ-Fe2O3@CR, and GG/γ-Fe2O3@NB when compared to the values for virgin GG/γ-Fe2O3 (Fig. S7) indicated the prevalence of variable covalent/ionic/hydrogen bonding interactions, owing to involvement of O–H, C[double bond, length as m-dash]O, etc. groups in bonding which influence the adsorption of CR and NB.62 For instance, the peak at 3378 cm−1 in the GG/γ-Fe2O3 shifted to a lower wavenumber, 3368 cm−1 in the GG/γ-Fe2O3@CR and 3354 cm−1 in GG/γ-Fe2O3@NB, due to extended hydrogen bonding.27 Moreover, the peak at 1622 cm−1 was assigned to O–H bending vibration in GG/γ-Fe2O3 alongside N[double bond, length as m-dash]N stretching of azo bonds in CR.118 However, no significant change in C–H stretching frequencies was discerned, and an additional peak at 1231 cm−1 was ascertained for GG/γ-Fe2O3@CR for S[double bond, length as m-dash]O stretch due to the –SO3 group in CR.119 The strong molecular bonding interactions between γ-Fe2O3 and functional groups on GG (–OH, –CO–, –COOH), as indicated by FT-IR spectra and XPS analysis, contributed to its stability and prevented any leaching possibility of the composite in water during adsorption.62 Thus, the commendable adsorption performance of GG/γ-Fe2O3 for both CR and NB might be the consequence of FT-IR established extensive functional sites on the GG/γ-Fe2O3 surface, and their specific interaction through several interactive pathways. Furthermore, the presence of charge on dyes, and GG/γ-Fe2O3 surface over and above the ZPC120 might have also paved the way to additional weak physical, and non-specific bonding interactions, viz., van der Waals, n–π, anion–π, and π–π stacking in addition to previously confirmed electrostatic and hydrogen bonding interactions (Fig. 7). The strength of these forces was confirmed from the bonding energy pattern obtained from the Temkin and D–R isotherms, alongside the thermodynamic investigations.
image file: d5ra04982b-f7.tif
Fig. 7 Plausible mechanism for the adsorption of CR and NB by GG/γ-Fe2O3.

From a mechanistic viewpoint, analysis of steps governing the adsorption process becomes an indispensable task. A detailed exploration suggested that the adsorption process could be controlled by either a mass action mechanism or a chemical action (IPD/LFD). The former was realised as irrelevant considering physisorption to be a fast phenomenon.121 Thus, the kinetic data was fitted to the Webber–Morris and Boyd relationships to assert whether the diffusion mechanism underlying physisorption for CR and NB removal by GG/γ-Fe2O3 followed (a) intraparticle diffusion (IPD), or (b) liquid film diffusion (LFD) kinetics, or (c) a simultaneous combination of both processes covering all the interior and exterior surface pores of the adsorbent by the pollutant ions.


3.4.5.1. Intraparticle diffusion. The Intraparticle diffusion (IPD) model is validated for those systems in which rapid adsorption takes place, such that the pollutant diffuses through the surface into the interstitial pores of the adsorbent, and binds through physical/chemical bonds, which is characterized as the RDS. This phenomenon is well understood by fitting sorption data into the Weber and Morris equation122 [eqn (S4)].

The linear IPD plots, for both CR and NB, did not pass through the origin, which suggested that IPD was not the sole RDS (Fig. S8a and b).123 Moreover, the plots depicted three linear regions for both dyes, the initial portion incorporated the diffusion of dyes into the exterior surface-active sites, followed by their gradual intrusion into the interstitial pores which later slowed down following unavailability of sorptive sites and diminished dye concentration, subjugating a three-step mechanism, and thus the multilinearity.124 In addition, the value of intercept C (Table S8) for NB (9.178) was superior to that for CR (4.240), indicative of significant coverage at the boundary layer of GG/γ-Fe2O3 on account of driving diffusion of NB.


3.4.5.2. Liquid film diffusion. The Liquid film diffusion (LFD) model holds good for adsorption systems in which mass diffusion/crossing of the boundary layer by the external liquid film of pollutant from the bulk, formulated around the solid interface and the surface-active sites is established as the RDS. The phenomenon can be justified using Boyd's equation116 [eqn (S5)].

The LFD plots showed linearity for CR and NB sorption but deviated from the origin, yielding a non-zero intercept value, which clarified that the adsorption was not solely governed by the film diffusion kinetics (Fig. S8c and d). The kLFD and R2 values are given in Table S8.

The foregoing observations suggested a specific chemical interaction of CR and NB with functional sites on the GG/γ-Fe2O3, which was found in coincidence with those from the PSO kinetic model. Thereby, it can be concluded that both CR and NB adsorption on GG/γ-Fe2O3 surface was governed partially by intraparticle as well as film diffusion steps, supplementing bulk transport and adsorptive attachment.

3.5. Photodegradation of CR dye, kinetics, and degradation mechanism

The photodegradation of azo-group-containing CR using GG/γ-Fe2O3 catalyst was performed at 38 °C under sunlight. From the time-dependent UV-visible absorption spectra, the absorbance of the CR solutions decreased over time upon degradation (Fig. S9a). The rate of dye degradation was consistent, and nearly 50% of the CR was degraded within the first 80 min, while up to 92.7% degradation was observed during 160 min of irradiation. This high rate of degradation might be a result of a prominent band gap of the GG/γ-Fe2O3. The PFO kinetics plot (Fig. S9b) of ln Ct/C0 vs. time t (min) produced a straight line with rate constant 0.014 min−1 and R2 = 0.877, indicating obedience to the PFO degradation mechanism. In contrast, the adsorption of CR over the GG/γ-Fe2O3 surface followed PSO kinetics. Therefore, the complete removal of CR could be attributed to the simultaneous adsorption in the dark, followed by photocatalytic degradation under sunlight.

From the mechanistic notion, it can be impounded that on absorption of solar radiation of energy hν, more than the band gap energy, the electrons (e) are excited from valence band (VB) to conduction band (CB), with simultaneous accumulation of electrons in CB, and holes (h+) in VB, respectively.125 Fe3+ ions have been reported previously to supress the electro-hole recombination rate, because besides serving as active sites for adsorption and activation, the d-orbitals in Fe3+ sites can also enhance the charge transfer efficiency.126 The dissolved oxygen (O2) gets reduced by the photo-induced electrons into superoxide radical anion (O2˙). Moreover, the positively charged holes oxidize H2O to OH˙.127 Further, to gain complete insight of the photodegradation mechanism, the catalytic effect of active species was investigated through free radical trapping experiment (Fig. S10). The CR degradation involving GG/γ-Fe2O3 was considerably inhibited by scavengers in order: EDTA (h+) < BQ (O2˙) < IPA (OH˙). The results revealed that the CR percentage degradation was primarily influenced by OH˙ and O2˙ oxidative species, followed by the holes (h+). These highly reactive oxygen species (ROS) are powerful oxidants for complete mineralization of CR to simple degradation products, like CO2, H2O, NH4+, NO3, SO42−, and mineral acids (Fig. 8). Furthermore, Fe2O3 and its nanocomposites have been extensively reported as efficient photocatalysts for dye degradation.128–130


image file: d5ra04982b-f8.tif
Fig. 8 Proposed mechanism for photodegradation of azo-CR by GG/γ-Fe2O3.

In order to determine the possible degradation intermediates, a time-based LC-MS analysis (Fig. S11) of photodegraded diazo-CR dye solution, at optimum experimental conditions, was examined. The major intermediate species formed during CR degradation, detected by LC-MS, are shown in the Scheme 2. During the initial analysis, partial cleavage of the azo (–N[double bond, length as m-dash]N–) bond or modification of side groups in CR might produce large high-mass aromatic fragments. Further, breakdown of large fragments resulted in complete azo bond scission, desulphonation and deamination following –C–S−/−C–N– bond cleavage, hydroxylation, azo reduction, oxidation, rearrangements, and cleavage of –C–C– bonds between the chromophore rings.5,131–135 This step produced various intermediates and their derivatives, including 4-aminonaphthalene-1-sulphonic acid (ANSA, m/z = 246.26), 4,4′-diaminobiphenyl (benzidine, m/z = 184.84), aniline-4-sulphonic acid (sulphanilic acid, m/z = 172.11), biphenyl (m/z = 150.99), 1-naphthylamine/2-naphthylamine (m/z = 141.13), aniline (m/z = 90.92), benzene-1-ylium (m/z = 77.05), benzene (m/z = 78), etc. Later, the oxidative ring opening steps formed low molecular weight aromatics, acid/amine intermediates, polyphenols, and ultimately mineralized into CO2, H2O, NH4+, NO3, and SO42−.


image file: d5ra04982b-s2.tif
Scheme 2 Proposed degradation pathways of CR by GG/γ-Fe2O3.

3.6. Real water analysis and competitive removal of CR and NB dyes

The adsorption/degradation tendency of the GG/γ-Fe2O3 for CR and NB amidst different water environments and co-existing ions is shown in Table S9. The removal efficiency hinders in natural wastewater samples due to the competition between large number of co-existing pollutants (organic, inorganic, and microorganisms) that complete for limited available surface-active sites. The % CR and NB removal decreased in tap water and sewage water, in comparison to RO water or distilled water, due to more competitive effect.

Typically, the solubility of organic contaminants decreases with the addition of salts (ions) owing to self-aggregation, i.e., salting-out or primary kinetic salt effect.136 This limits the dye solubility due to fewer available water molecules. However, the CR adsorption remained largely unaffected in the presence of salts. The results showed little effect of co-existing ions on CR adsorption (except that of CO32−), indicating excellent resistance of GG/γ-Fe2O3 to ion interference, and thus confirming specific CR adsorption by GG/γ-Fe2O3. The interference by CO32− might be mainly due its interaction with CR, and not salting.136 Moreover, the inhibitory effect of divalent CO32− (most) and SO42− on CR adsorption/degradation were greater than monovalent Cl or NO3, which can be attributed to strong electrostatic attractions between GG/γ-Fe2O3 and the higher anions, under similar conditions.137 Similar observations were drawn for NB adsorption. Additionally, CO32− ions highly and negatively affected percentage CR degradation, due to the hydrolysis of HCO3, an OH˙ scavenger.138 Moreover, percentage NB adsorption decreased around cations vis-à-vis anions, suggesting higher cationic interface potential for cationic NB adsorption surfacing weak electrostatic interactions. Previous studies have shown that both CR and NB undergo stable complexation with metal ions which can affect their adsorption/degradation onto support materials in presence of different ions.139 Furthermore, the competitive influence of other anionic and cationic dyes on the CR and NB removal by GG/γ-Fe2O3 from a mixture is shown in Table S10. The results showed the specificity and efficiency of GG/γ-Fe2O3 towards CR and NB adsorption, and substantial potential in wastewater treatment.

3.7. Regeneration and reusability of GG/γ-Fe2O3

Fig. S12 shows the reusability results of the GG/γ-Fe2O3 for CR and NB adsorption up to six cycles. The GG/γ-Fe2O3 was found efficient till six consecutive cycles for CR (97.12% to 76.27%), and up to 3 cycles for NB (97.12% to 71.28%). Therefore, the GG/γ-Fe2O3 can be more efficiently used for CR , i.e., anionic dye than NB, i.e., cationic dye, sorption for potential practical applications. Further, the FT-IR analysis of the spent GG/γ-Fe2O3 after CR adsorption (spent GG/γ-Fe2O3@CR) and NB adsorption (spent GG/γ-Fe2O3@NB), Fig. S7, exhibited comparable peak positions to pristine GG/γ-Fe2O3, establishing appreciable structural and functional stability.

3.8. Cost analysis of GG/γ-Fe2O3

The commercial and large-scale implementation of the GG/γ-Fe2O3 for dye remediation from industrial effluents widely depends on its cost-effectiveness. The study addresses detailed stepwise consideration of economic feasibility through (a) the cost of synthesis of GG/γ-Fe2O3 per batch (on laboratory scale) (Table S11), and (b) their processing cost for use as an adsorbent/catalyst for treating 1000 L of wastewater using GG/γ-Fe2O3 (Table S12). The synthetic one-pot co-precipitation approach is comparatively simpler and cost-efficient in comparison to other sophisticated methods.44 The Glycyrrhiza glabra roots were locally sourced at negligible cost, and were not subjected to any sort of energy-intensive chemical pre-treatment or post synthetic modification through expensive surfactants, toxic solvents, reagents/chemicals or synthetic stabilizers for a greener approach. Additionally, the energy consumption, limited to heating/drying, magnetic stirring, and calcination was moderate, scalable for industrial level production. The lab scale production of the GG/γ-Fe2O3 was estimated around $0.79 to $0.88, which is significantly lower than the previously reported commercial materials, including silver nanoparticles ($15.77) and nanopowders ($20.42) for similar applications.140,141 The majority of synthesis expense was due to analytical lab-grade γ-Fe2O3 precursors of high purity, which can be drastically reduced by up to 100 times on large scale using industrial-grade reagents. From the perspective of treatment cost, $158 to $176 was estimated to treat 1000 L wastewater using optimum dosage 0.2 g L−1 of GG/γ-Fe2O3, considering unit synthesis cost of GG/γ-Fe2O3 for adsorption. Solar-photocatalysis was employed for dye degradation, which cuts down the power consumption. Considering practically feasible regeneration up to four cycles, the effective treatment cost reduced by 70–80%, in range $31.6 to $52.8 per 1000 L of wastewater.

3.9. Comparison study

The comparative evaluation of GG/γ-Fe2O3 adsorbent/photocatalyst's efficiency in CR and NB dyes remediation was estimated with reference to the available literature (Tables S13 and S14). The greater effectiveness of the GG/γ-Fe2O3, in comparison to previously studied materials, established its wider applicability. This validation, supported by cost-effectiveness due to simple synthetic approach, and implementation of highly accessible precursors accounts for potential reliability in realistic and practical scenario.

4 Conclusions and future prospects

In summary, a structurally, thermally, and functionally stable GG/γ-Fe2O3 nanocomposite was synthesized in situ via simple co-precipitation method, and employed for subsequent adsorption and photodegradation of CR and NB dyes, with pursuit of superior antimicrobial activity compared to the parent precursor GG. Both CR and NB dyes are extremely hazardous water pollutants that find application in various industries. The GG/γ-Fe2O3 facilitated appreciable sorption capacity (Qo) of 47.504 mg g−1, and 15.361 mg g−1 for CR and NB at 50 and 30 °C, respectively, achieving 99.78% and 97.23% decolorization of CR and NB, respectively, through adsorption under optimized conditions. The nanocomposite provided 92.7% degradation of CR in 160 min with a pseudo-first order rate constant of 0.014 min−1, dominated by OH˙ and O2˙ active species. The applicability of the GG/γ-Fe2O3 for real wastewater treatment amongst competing ions and dye mixtures, under the same set of optimized experimental conditions illustrate the reliability and transferability of these conditions for large-scale application, while achieving balance between selectivity and efficiency. The comparative economic and evaluative performance of the GG/γ-Fe2O3 with previously synthesized materials, accompanied by regenerative tendency and FT-IR affirmed stability of spent GG/γ-Fe2O3 post treatment indicated significant potential in water treatment. Overall, the study bridges laboratory outcomes with practical applications and contributes to the development of advanced multifunctional materials for decolourising dye-laden industrial wastewater.

However, a proper post-treatment disposal strategy, for the composite, should be designed considering environmental sustainability. Additionally, changes in total organic carbon (TOC) and water quality parameters should be investigated to determine the mineralization efficiency after dye degradation. The future work may address the integration of GG/γ-Fe2O3 for continuous flow or field-scale treatment systems.

Author contributions

Ankita Manchanda: conceptualization, visualisation, formal analysis, investigation, methodology, software, data curation, writing-original draft, funding acquisition. Ahmed Hussain Jawhari: data curation, validation. Ziaul Hasan: conceptualization, investigation. Nazim Hasan: data curation, validation. Sneha Shukla: validation. Adiba Khan: validation. Tabrez Alam Khan: supervision, writing-review & editing. Saif Ali Chaudhry: supervision, project administration, resources, writing-review & editing.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

The authors confirm that the data supporting the findings of the study are available within the article and in its supplementary information (SI). Supplementary information is available. See DOI: https://doi.org/10.1039/d5ra04982b.

Acknowledgements

The authors highly acknowledge Jamia Millia Islamia, New Delhi, for providing laboratory facilities. Ankita Manchanda expresses her gratitude to the Department of Science and Technology (DST), New Delhi, India, for financial support through INSPIRE fellowship (IF200064), and Saif Ali Chaudhry acknowledges financial support by DST-SERB (EEQ/2022/001063).

References

  1. D. Dutta, S. Arya and S. Kumar, Industrial wastewater treatment: current trends, bottlenecks, and best practices, Chemosphere, 2021, 285, 131245 CrossRef PubMed.
  2. W. A. Shaikh, R. U. Islam and S. Chakraborty, Stable silver nanoparticle doped mesoporous biochar-based nanocomposite for efficient removal of toxic dyes, J. Environ. Chem. Eng., 2021, 9, 104982 CrossRef.
  3. Top Dyes and Pigments Companies in Global [Updated]|Global Growth Insights, https://www.globalgrowthinsights.com/blog/dyes-and-pigments-companies-733, accessed 18 June 2025 Search PubMed.
  4. S. Moosavi, C. W. Lai, S. Gan, G. Zamiri, O. A. Pivehzhani and M. R. Johan, Application of efficient magnetic particles and activated carbon for dye removal from wastewater, ACS Omega, 2020, 5, 20684–20697 CrossRef PubMed.
  5. P. Yadav, N. Dhariwal, M. Kumari, V. Kumar and O. P. Thakur, Enhanced degradation of Congo-red dye by Cr3+ doped α-Fe2O3 nano-particles under sunlight and industrial wastewater treatment, Chemosphere, 2023, 343, 140208 CrossRef PubMed.
  6. S. Meneceur, H. Hemmami, A. Bouafia, S. E. Laouini, M. L. Tedjani, D. Berra and M. S. Mahboub, Photocatalytic activity of iron oxide nanoparticles synthesized by different plant extracts for the degradation of diazo dyes Evans blue and Congo red, Biomass Convers. Biorefin., 2024, 14, 5357–5372 CrossRef.
  7. G. Sarojini, S. V. Babu and M. Rajasimman, Adsorptive potential of iron oxide-based nanocomposite for the sequestration of Congo red from aqueous solution, Chemosphere, 2022, 287, 132371 CrossRef PubMed.
  8. N. Natasha, A. Khan, U. U. Rahman, N. Sadaf, M. Yaseen, R. A. Abumousa, R. Khattak, N. Rehman, M. Bououdina and M. Humayun, Effective Removal of Nile Blue Dye from Wastewater using Silver-Decorated Reduced Graphene Oxide, ACS Omega, 2024, 9, 19461–19480 CrossRef PubMed.
  9. E. Jafarian, A. Hekmatiyan, A. Cheraghdar, H. Safarzadeh and M. Shamsi, Elimination performance of Nile blue from wastewater using by carboxymethyl cellulose-graft-poly(methacrylic acid-co-acrylamide)/kaolin nanocomposite hydrogel, Int. J. Environ. Sci. Technol., 2023, 20, 9933–9944 CrossRef.
  10. A. K. Nayak and A. Pal, Statistical modeling and performance evaluation of biosorptive removal of Nile blue A by lignocellulosic agricultural waste under the application of high-strength dye concentrations, J. Environ. Chem. Eng., 2020, 8, 103677 CrossRef.
  11. D. Hussain, S. A. Khan, T. A. Khan and S. S. Alharthi, Efficient liquid phase confiscation of nile blue using a novel hybrid nanocomposite synthesized from guar gum-polyacrylamide and erbium oxide, Sci. Rep., 2022, 12, 14656 CrossRef PubMed.
  12. A. Iqbal, A. Yusaf, M. Usman, T. H. Bokhari and A. Mansha, Insight into the degradation of different classes of dyes by advanced oxidation processes; a detailed review, Int. J. Environ. Anal. Chem., 2023, 5503–5537 Search PubMed.
  13. R. Gusain, K. Gupta, P. Joshi and O. P. Khatri, Adsorptive removal and photocatalytic degradation of organic pollutants using metal oxides and their composites: a comprehensive review, Adv. Colloid Interface Sci., 2019, 272, 102009 Search PubMed.
  14. B. Lellis, C. Z. Fávaro-Polonio, J. A. Pamphile and J. C. Polonio, Effects of textile dyes on health and the environment and bioremediation potential of living organisms, Biotechnol. Res. Innov., 2019, 3, 275–290 Search PubMed.
  15. M. A. Alkhadra, X. Su, M. E. Suss, H. Tian, E. N. Guyes, A. N. Shocron, K. M. Conforti, J. P. De Souza, N. Kim, M. Tedesco, K. Khoiruddin, I. G. Wenten, J. G. Santiago, T. A. Hatton and M. Z. Bazant, Electrochemical Methods for Water Purification, Ion Separations, and Energy Conversion, Chem. Rev., 2022, 122, 13547–13635 Search PubMed.
  16. P. Ganguly, R. Sarkhel and P. Das, Synthesis of pyrolyzed biochar and its application for dye removal: batch, kinetic and isotherm with linear and non-linear mathematical analysis, Surf. Interfaces, 2020, 20, 100616 Search PubMed.
  17. M. Wen, G. Li, H. Liu, J. Chen, T. An and H. Yamashita, Metal-organic framework-based nanomaterials for adsorption and photocatalytic degradation of gaseous pollutants: recent progress and challenges, Environ. Sci. Nano, 2019, 6, 1006–1025 Search PubMed.
  18. S. Nangia, D. Katyal and S. G. Warkar, Thermodynamics, kinetics and isotherm studies on the removal of anionic azo-dye (Congo red) using synthesized chitosan/Moringa oleifera gum hydrogel composites, Sep. Sci. Technol., 2023, 58, 13–28 Search PubMed.
  19. S. A. Khan, A. Manchanda and T. A. Khan, Adsorption of Coomassie Brilliant Blue on a Novel Eco-Friendly Nanogel from Simulated Water: Equilibrium, Kinetic, and Thermodynamic Studies, ChemistrySelect, 2024, 9, e202304927 Search PubMed.
  20. M. Yaseen, M. Humayun, A. Khan, M. Idrees, N. Shah and S. Bibi, Photo-Assisted Removal of Rhodamine B and Nile Blue Dyes from Water Using CuO–SiO2 Composite, Molecules, 2022, 27(16), 5343 Search PubMed.
  21. Y. Pi, C. Duan, Y. Zhou, S. Sun, Z. Yin, H. Zhang, C. Liu and Y. Zhao, The effective removal of Congo Red using a bio-nanocluster: Fe3O4 nanoclusters modified bacteria, J. Hazard. Mater., 2022, 424, 127577 CrossRef PubMed.
  22. S. I. Siddiqui and S. A. Chaudhry, Nanohybrid composite Fe2O3-ZrO2/BC for inhibiting the growth of bacteria and adsorptive removal of arsenic and dyes from water, J. Clean. Prod., 2019, 223, 849–868 CrossRef.
  23. M. Kasbaji, I. Ibrahim, M. Mennani, O. a. abuelalla, S. S. fekry, M. M. Mohamed, T. M. Salama, I. A. Moneam, M. Mbarki, A. Moubarik and M. Oubenali, Future trends in dye removal by metal oxides and their nano/composites: a comprehensive review, Inorg. Chem. Commun., 2023, 158, 111546 CrossRef.
  24. A. Tchinsa, M. F. Hossain, T. Wang and Y. Zhou, Removal of organic pollutants from aqueous solution using metal organic frameworks (MOFs)-based adsorbents: a review, Chemosphere, 2021, 284, 131393 CrossRef PubMed.
  25. K. Manzoor, M. Batool, F. Naz, M. F. Nazar, B. H. Hameed and M. N. Zafar, A comprehensive review on application of plant-based bioadsorbents for Congo red removal, Biomass Convers. Biorefin., 2024, 14, 4511–4537 CrossRef.
  26. N. Kumar, A. Pandey, Rosy and Y. C. Sharma, A review on sustainable mesoporous activated carbon as adsorbent for efficient removal of hazardous dyes from industrial wastewater, J. Water Process Eng., 2023, 54, 104054 CrossRef.
  27. L. Cheng, Y. Ji, X. Liu, L. Mu and J. Zhu, Sorption mechanism of organic dyes on a novel self-nitrogen-doped porous graphite biochar: coupling DFT calculations with experiments, Chem. Eng. Sci., 2021, 242, 116739 CrossRef.
  28. K. K. Kefeni and B. B. Mamba, Photocatalytic application of spinel ferrite nanoparticles and nanocomposites in wastewater treatment: review, Sustain. Mater. Technol., 2020, 23, e00140 Search PubMed.
  29. G. Fadillah, S. P. Yudha, S. Sagadevan, I. Fatimah and O. Muraza, Magnetic iron oxide/clay nanocomposites for adsorption and catalytic oxidation in water treatment applications, Open Chem., 2020, 18, 1148–1166 Search PubMed.
  30. J. Wang, Y. Chen, G. Liu and Y. Cao, Synthesis, characterization and photocatalytic activity of inexpensive and non-toxic Fe2O3−Fe3O4 nano-composites supported by montmorillonite and modified by graphene, Composites, Part B, 2017, 114, 211–222 CrossRef.
  31. X. Wang, L. Tang, T. Yang, Y. Shi, F. Liu and H. Jiang, Remediation of bensulfuron methyl polluted water and soil by Fe2O3/Fe3O4@C activating peroxymonosulfate: chloride enhancement effect and phytotoxicity assessment, Chem. Eng. J., 2023, 474, 145439 CrossRef.
  32. S. Yadav, K. Shakya, A. Gupta, D. Singh, A. R. Chandran, A. V. Aanappalli, K. Goyal, N. Rani and K. Saini, A review on degradation of organic dyes by using metal oxide semiconductors, Environ. Sci. Pollut. Res., 2023, 30, 71912–71932 CrossRef PubMed.
  33. S. Arsalani, E. J. Guidelli, J. F. D. F. Araujo, A. C. Bruno and O. Baffa, Green Synthesis and Surface Modification of Iron Oxide Nanoparticles with Enhanced Magnetization Using Natural Rubber Latex, ACS Sustain. Chem. Eng., 2018, 6, 13756–13765 CrossRef.
  34. S. Mishra, L. Cheng and A. Maiti, The utilization of agro-biomass/byproducts for effective bio-removal of dyes from dyeing wastewater: a comprehensive review, J. Environ. Chem. Eng., 2021, 9, 104901 CrossRef.
  35. N. J. Ashbolt, Microbial Contamination of Drinking Water and Human Health from Community Water Systems, Curr. Environ. Health Rep., 2015, 2, 95–106 CrossRef PubMed.
  36. N. A. Salahuddin, H. A. EL-Daly, R. G. El Sharkawy and B. T. Nasr, Nano-hybrid based on polypyrrole/chitosan/graphene oxide magnetite decoration for dual function in water remediation and its application to form fashionable colored product, Adv. Powder Technol., 2020, 31, 1587–1596 CrossRef.
  37. K. Naseem, M. H. Tahir, F. Farooqi, S. Manzoor and S. U. Khan, Strategies adopted for the preparation of sodium alginate–based nanocomposites and their role as catalytic, antibacterial, and antifungal agents, Rev. Chem. Eng., 2023, 39, 1359–1391 Search PubMed.
  38. U. Baig, M. Faizan and M. Sajid, Effective removal of hazardous pollutants from water and deactivation of water-borne pathogens using multifunctional synthetic adsorbent materials: a review, J. Clean. Prod., 2021, 302, 126735 Search PubMed.
  39. T. K. Lim, Glycyrrhiza glabra, Edible Medicinal and Non-medicinal Plants, 2016, pp. 354–457 Search PubMed.
  40. E. Alibakhshi, M. Ramezanzadeh, S. A. Haddadi, G. Bahlakeh, B. Ramezanzadeh and M. Mahdavian, Persian liquorice extract as a highly efficient sustainable corrosion inhibitor for mild steel in sodium chloride solution, J. Clean. Prod., 2019, 210, 660–672 CrossRef.
  41. A. Khan, Z. Sheerazi, S. Shukla, A. Manchanda and S. A. Chaudhry, Multifunctional nanocomposite for sustainable Yamuna water treatment: insight into adsorption, statistical optimization, and antibacterial action, Chemosphere, 2025, 385, 144565 CrossRef PubMed.
  42. M. Rui, C. Ma, Y. Hao, J. Guo, Y. Rui, X. Tang, Q. Zhao, X. Fan, Z. Zhang, T. Hou and S. Zhu, Iron oxide nanoparticles as a potential iron fertilizer for peanut (Arachis hypogaea), Front. Plant Sci., 2016, 7, 195361 Search PubMed.
  43. A. Saurabh, M. Kaur, R. Khan, G. Guleria, M. Shandilya and S. Thakur, Foliar application of Fe2O3 nanofertilizer on growth and yield of cauliflower (Brassica oleracea var. Botrytis L.) cv. Pusa Snowball K-1, Int. J. Phytoremediation, 2024, 26, 993–1002 CrossRef PubMed.
  44. A. Herawati, R. M. Aryani, G. Antarnusa, Khoiriah and A. Nene, Fe3O4/chitosan nanocomposites for Congo red removal: adsorption surpasses photodegradation, Mater. Chem. Phys., 2026, 347, 131431 CrossRef.
  45. T. Shahriari, H. Kangazian and B. A. Goharrizi, Application of Licorice-derived Activated Carbon in Tannery Wastewater Treatment, J. Med. Plants By-Prod., 2024, 13, 741–746 Search PubMed.
  46. S. Z. Mohammadi, H. Hamidian and Z. Moeinadini, High surface area-activated carbon from Glycyrrhiza glabra residue by ZnCl2 activation for removal of Pb(II) and Ni(II) from water samples, J. Ind. Eng. Chem., 2014, 20, 4112–4118 CrossRef.
  47. B. Ovez, S. Ozgen and M. Yuksel, Biological denitrification in drinking water using Glycyrrhiza glabra and Arunda donax as the carbon source, Process Biochem., 2006, 41, 1539–1544 CrossRef.
  48. S. Pirsalami, S. Bagherpour, M. E. Bahrololoom and M. Riazi, Adsorption efficiency of glycyrrhiza glabra root toward heavy metal ions: experimental and molecular dynamics simulation study on removing copper ions from wastewater, Sep. Purif. Technol., 2021, 275, 119215 CrossRef.
  49. A. Ramos, E. Monteiro and A. Rouboa, Biomass pre-treatment techniques for the production of biofuels using thermal conversion methods – a review, Energy Convers. Manag., 2022, 270, 116271 CrossRef.
  50. A. Choudhry, A. Sharma, T. A. Khan and S. A. Chaudhry, Flax seeds based magnetic hybrid nanocomposite: an advance and sustainable material for water cleansing, J. Water Process Eng., 2021, 42, 102150 CrossRef.
  51. M. Mahmoodi, B. Aslibeiki, S. Ghosh, L. Hasani, S. Slimani, L. Vattuone, D. Peddis and T. Sarkar, Apple Tree Root-Derived Biochar/Iron Oxide Triphasic Nanocomposite for Wastewater Treatment and Microwave Absorption, Adv. Sustainable Syst., 2024, 9, 2400549 CrossRef.
  52. G. Sposito, On Points of Zero Charge, Environ. Sci. Technol., 1998, 32, 2815–2819 CrossRef.
  53. A. Choudhry, A. Sharma, S. I. Siddiqui, I. Ahamad, M. Sajid, T. A. Khan and S. A. Chaudhry, Origanum vulgare manganese ferrite nanocomposite: an advanced multifunctional hybrid material for dye remediation, Environ. Res., 2023, 220, 115193 CrossRef PubMed.
  54. J. H. Jorgensen and M. J. Ferraro, Antimicrobial susceptibility testing: a review of general principles and contemporary practices, Clin. Infect. Dis., 2009, 49, 1749–1755 CrossRef.
  55. M. Irani, M. Sarmadi, F. Bernard, H. Ebrahimi and H. S. Bazarnov, Leaves Antimicrobial Activity of Glycyrrhiza glabra L, Iran. J. Pharm. Res., 2010, 9(4), 425–82010 Search PubMed.
  56. Z. Hasan, A. Islam and L. A. Khan, Signature Garlic Phytochemical as a Potential Anti-Candidal Candidate Targeting Virulence Factors in Candida albicans, Med. Sci. Forum., 2023, 21(1), 50 Search PubMed.
  57. W. PA, Reference Method for Broth Dilution Antifungal Susceptibility Testing of Yeasts, Approved Standard-second edition, CLSI document M27-A2, 2002 Search PubMed.
  58. M. D. Olawale, E. O. Akintemi, B. E. Agbaffa and J. A. Obaleye, Synthesis, characterization, adsorption study, quantum mechanics, monte carlo and molecular dynamics of lead based polymeric compound towards mopping of aqueous methyl red dye, Results Chem., 2022, 4, 100499 CrossRef.
  59. T. A. Khan, M. Nazir and E. A. Khan, Magnetically modified multiwalled carbon nanotubes for the adsorption of bismarck brown R and Cd(II) from aqueous solution: batch and column studies, Desalination Water Treat., 2016, 57, 19374–19390 CrossRef CAS.
  60. I. Ahmad, K. Manzoor, G. Aalam, M. Amir, S. W. Ali and S. Ikram, Facile Synthesis of L-Tryptophan Functionalized Magnetic Nanophotocatalyst Supported by Copper Nanoparticles for Selective Reduction of Organic Pollutants and Degradation of Azo Dyes, Catal. Lett., 2023, 153, 2604–2623 CrossRef CAS.
  61. J. T. Schneider, D. S. Firak, R. R. Ribeiro and P. Peralta-Zamora, Use of scavenger agents in heterogeneous photocatalysis: truths, half-truths, and misinterpretations, Phys. Chem. Chem. Phys., 2020, 22, 15723–15733 RSC.
  62. H. Wang, Q. Yao, C. Wang, B. Fan, Q. Sun, C. Jin, Y. Xiong and Y. Chen, A simple, one-step hydrothermal approach to durable and robust superparamagnetic, superhydrophobic and electromagnetic wave-absorbing wood, Sci. Rep., 2016, 6, 1–10 CrossRef CAS.
  63. A. Rifianto, J. Widakdo, N. Istikhomah, E. Suharyadi, T. Kato and S. Iwata, The effect of synthesis parameter on crystal structure and magnetic properties of Ni0.5Zn0.5Fe2O4 magnetic nanoparticles, J. Phys. Conf., 2018, 1011, 012059 CrossRef.
  64. L. Rahman, S. Bhattacharjee, S. Islam, F. Zahan, B. Biswas and N. Sharmin, A study on the preparation and characterization of maghemite (γ-Fe2O3) particles from iron-containing waste materials, J. Asian Ceram. Soc., 2020, 8, 1083–1094 CrossRef.
  65. M. M. Ba-Abbad, A. Benamour, D. Ewis, A. W. Mohammad and E. Mahmoudi, Synthesis of Fe3O4 Nanoparticles with Different Shapes Through a Co-Precipitation Method and Their Application, J. Met., 2022, 74, 3531–3539 CAS.
  66. F. Yang, T. Chu, Y. Zhang, X. Liu, G. Sun and Z. Chen, Quality assessment of licorice (Glycyrrhiza glabra L.) from different sources by multiple fingerprint profiles combined with quantitative analysis, antioxidant activity and chemometric methods, Food Chem., 2020, 324, 126854 CrossRef CAS PubMed.
  67. M. Bavanilatha, L. Yoshitha, S. Nivedhitha and S. Sahithya, Bioactive studies of TiO2 nanoparticles synthesized using Glycyrrhiza glabra, Biocatal. Agric. Biotechnol., 2019, 19, 101131 CrossRef.
  68. A. Yousefi, S. A. S. Ebrahimi, A. Seyfoori and H. M. Hosseini, Maghemite Nanorods and Nanospheres: Synthesis and Comparative Physical and Biological Properties, Bionanoscience, 2018, 8, 95–104 CrossRef.
  69. Y. Huo, P. Singh, Y. J. Kim, V. Soshnikova, J. Kang, J. Markus, S. Ahn, V. Castro-Aceituno, R. Mathiyalagan, M. Chokkalingam, K. S. Bae and D. C. Yang, Biological synthesis of gold and silver chloride nanoparticles by Glycyrrhiza uralensis and in vitro applications, Artif. Cells, Nanomed., Biotechnol., 2018, 46, 303–312 CrossRef CAS PubMed.
  70. P. Pascuta, G. Borodi, M. Bosca, L. Pop, S. Rada and E. Culea, Preparation and structural characterization of some Fe2O3-B2O3-ZnO glasses and glass ceramics, J. Phys.:Conf. Ser., 2009, 182, 012072 CrossRef.
  71. D. Liu, Z. Ma, Z. Wang, H. Tian and M. Gu, Biodegradable poly(vinyl alcohol) foams supported by cellulose nanofibrils: processing, structure, and properties, Langmuir, 2014, 30, 9544–9550 CrossRef CAS.
  72. M. E. Mejia-Santillan, N. Pariona, J. Bravo-C., M. Herrera-Trejo, F. Montejo-Alvaro, A. Zarate, D. L. Perry and A. I. Mtz-Enriquez, Physical and arsenic adsorption properties of maghemite and magnetite sub-microparticles, J. Magn. Magn. Mater., 2018, 451, 594–601 CrossRef CAS.
  73. M. Arakha, S. Pal, D. Samantarrai, T. K. Panigrahi, B. C. Mallick, K. Pramanik, B. Mallick and S. Jha, Antimicrobial activity of iron oxide nanoparticle upon modulation of nanoparticle-bacteria interface, Sci. Rep., 2015, 5, 1–12 Search PubMed.
  74. M. Khatamian, N. A. No, S. H. Nami and S. Fazli-Shokouhi, Synthesis and characterization of zeolite A, Fe3O4/zeolite A, and Fe2O3/zeolite A nanocomposites and investigation of their arsenic removal performance, J. Iran. Chem. Soc., 2023, 20, 1657–1670 CrossRef CAS.
  75. M. Thommes, K. Kaneko, A. V. Neimark, J. P. Olivier, F. Rodriguez-Reinoso, J. Rouquerol and K. S. W. Sing, Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report), Pure Appl. Chem., 2015, 87, 1051–1069 CrossRef CAS.
  76. N. Srivastava, M. Srivastava, A. Alhazmi, A. Mohammad, S. Khan, D. B. Pal, S. Haque, R. Singh, P. K. Mishra and V. K. Gupta, Sustainable green approach to synthesize Fe3O4/α-Fe2O3 nanocomposite using waste pulp of Syzygium cumini and its application in functional stability of microbial cellulases, Sci. Rep., 2021, 11, 1–12 CrossRef PubMed.
  77. M. Roslova, Z. Huang and X. Zou, Structural studies of inorganic materials by electron crystallography, Comprehensive Inorganic Chemistry III, 3rd edn, 2023, vol. 1–10, pp. 51–85 Search PubMed.
  78. Z. K. Heiba, S. I. Ahmed and M. B. Mohamed, Improved nonlinear optical and magnetic properties of Eu-doped nano-maghemite, Appl. Phys. A:Mater. Sci. Process., 2022, 128, 1–11 CrossRef.
  79. R. Mahajan, S. Suriyanarayanan and I. A. Nicholls, Improved solvothermal synthesis of γ-Fe2O3 magnetic nanoparticles for SiO2 coating, Nanomaterials, 2021, 11, 1889 CrossRef CAS PubMed.
  80. Y. Li, C. Zhu, T. Lu, Z. Guo, D. Zhang, J. Ma and S. Zhu, Simple fabrication of a Fe2O3/carbon composite for use in a high-performance lithium ion battery, Carbon, 2013, 52, 565–573 CAS.
  81. J. P. Jyothibasu, R. H. Wang, K. Ong, J. H. L. Ong and R. H. Lee, Scalable synthesis of γ-Fe2O3–based composite films as freestanding negative electrodes with ultra-high areal capacitances for high-performance asymmetric supercapacitors, Cellulose, 2022, 29, 321–340 CAS.
  82. S. Sun, M. Sun, Y. Kong, Y. Fang and Y. Yao, MoS2 and graphene as dual, cocatalysts for enhanced visible light photocatalytic activity of Fe2O3, J. Sol. Gel Sci. Technol., 2016, 80, 719–727 CrossRef CAS.
  83. Y. F. Huang, W. H. Kuan, P. T. Chiueh and S. L. Lo, Pyrolysis of biomass by thermal analysis–mass spectrometry (TA–MS), Bioresour. Technol., 2011, 102, 3527–3534 CrossRef CAS PubMed.
  84. F. Batool, S. Kanwal, H. Kanwal, S. Noreen, M. S. Hodhod, M. Mustaqeem, G. Sharif, H. K. Naeem, J. Zahid and A. R. Z. Gaafar, Ecofriendly Synthesis of Magnetic Composites Loaded on Rice Husks for Acid Blue 25 Decontamination: Adsorption Kinetics, Thermodynamics, and Isotherms, Molecules, 2023, 28, 7124 CrossRef CAS PubMed.
  85. S. Ashrafi-Saiedlou, M. H. Rasouli-Sadaghiani, M. Fattahi and Y. Ghosta, Biosynthesis and characterization of iron oxide nanoparticles fabricated using cell-free supernatant of Pseudomonas fluorescens for antibacterial, antifungal, antioxidant, and photocatalytic applications, Sci. Rep., 2025, 15, 1–22 CrossRef.
  86. M. Shaban, M. Binsabt, A. M. Ahmed and F. Mohamed, Recycling rusty iron with natural zeolite heulandite to create a unique nanocatalyst for green hydrogen production, Nanomaterials, 2021, 11, 3445 CrossRef CAS PubMed.
  87. S. Saleem, M. N. Ashiq, S. Manzoor, U. Ali, R. Liaqat, A. Algahtani, S. Mujtaba, V. Tirth, A. M. Alsuhaibani, M. S. Refat, A. Ali, M. Aslam and A. Zaman, Analysis and characterization of opto-electronic properties of iron oxide (Fe2O3) with transition metals (Co, Ni) for the use in the photodetector application, J. Mater. Res. Technol., 2023, 25, 6150–6166 CrossRef CAS.
  88. P. Kumar, U. Kumar, Y. C. Huang, P. Y. Tsai, C. H. Liu, C. H. Wu, W. M. Huang and K. L. Chen, Photocatalytic activity of a hydrothermally synthesized γ-Fe2O3@Au/MoS2 heterostructure for organic dye degradation under green light, J. Photochem. Photobiol., A, 2022, 433, 114186 CrossRef CAS.
  89. N. Zahmouli, M. Hjiri, L. El Mir, A. Bonavita, N. Donato, G. Neri and S. G. Leonardi, High performance acetone sensor based on γ-Fe2O3/Al–ZnO nanocomposites, Nanotechnology, 2018, 30, 055502 CrossRef PubMed.
  90. V. K. Gupta, A. Fatima, U. Faridi, A. S. Negi, K. Shanker, J. K. Kumar, N. Rahuja, S. Luqman, B. S. Sisodia, D. Saikia, M. P. Darokar and S. P. S. Khanuja, Antimicrobial potential of Glycyrrhiza glabra roots, J. Ethnopharmacol., 2008, 116, 377–380 CrossRef PubMed.
  91. S. Rodino, A. Butu, M. Butu, P. Cornea and B. Marasti, Comparative studies on antibacterial activity of licorice, elderberry and dandelion, Dig. J. Nanomater. Biostruct., 2015, 10(3), 947–955 Search PubMed.
  92. N. T. T. Nguyen, L. M. Nguyen, T. T. T. Nguyen, T. T. Nguyen, D. T. C. Nguyen and T. Van Tran, Formation, antimicrobial activity, and biomedical performance of plant-based nanoparticles: a review, Environ. Chem. Lett., 2022, 20, 2531–2571 CrossRef CAS PubMed.
  93. G. Sharma, M. Naushad, A. Kumar, S. Rana, S. Sharma, A. Bhatnagar, F. J. Stadler, A. A. Ghfar and M. R. Khan, Efficient removal of coomassie brilliant blue R-250 dye using starch/poly(alginic acid-cl-acrylamide) nanohydrogel, Process Saf. Environ. Prot., 2017, 109, 301–310 CrossRef CAS.
  94. R. Nodehi, H. Shayesteh and A. R. Kelishami, Enhanced adsorption of congo red using cationic surfactant functionalized zeolite particles, Microchem. J., 2020, 153, 104281 CrossRef CAS.
  95. C. Su, Environmental implications and applications of engineered nanoscale magnetite and its hybrid nanocomposites: a review of recent literature, J. Hazard. Mater., 2017, 322, 48–84 CrossRef CAS PubMed.
  96. J. Iqbal, N. S. Shah, M. Sayed, N. K. Niazi, M. Imran, J. A. Khan, Z. U. H. Khan, A. G. S. Hussien, K. Polychronopoulou and F. Howari, Nano-zerovalent manganese/biochar composite for the adsorptive and oxidative removal of Congo-red dye from aqueous solutions, J. Hazard. Mater., 2021, 403, 123854 CrossRef CAS PubMed.
  97. E. Rápó and S. Tonk, Factors Affecting Synthetic Dye Adsorption; Desorption Studies: A Review of Results from the Last Five Years (2017–2021), Molecules, 2021, 26, 5419 CrossRef PubMed.
  98. X. C. Tang, M. J. Pikal and L. S. Taylor, The Effect of Temperature on Hydrogen Bonding in Crystalline and Amorphous Phases in Dihydropyrine Calcium Channel Blockers, Pharm. Res., 2002, 19, 484–490 CrossRef CAS.
  99. M. C. S. Reddy, L. Sivaramakrishna and A. V. Reddy, The use of an agricultural waste material, Jujuba seeds for the removal of anionic dye (Congo red) from aqueous medium, J. Hazard. Mater., 2012, 203–204, 118–127 CrossRef PubMed.
  100. A. Haleem, A. Shafiq, S. Q. Chen and M. Nazar, A Comprehensive Review on Adsorption, Photocatalytic and Chemical Degradation of Dyes and Nitro-Compounds over Different Kinds of Porous and Composite Materials, Molecules, 2023, 28(3), 1081 CrossRef CAS PubMed.
  101. J. Nasiri, E. Motamedi, M. R. Naghavi and M. Ghafoori, Removal of crystal violet from water using β-cyclodextrin functionalized biogenic zero-valent iron nanoadsorbents synthesized via aqueous root extracts of Ferula persica, J Hazard. Mater., 2019, 367, 325–338 CrossRef CAS PubMed.
  102. S. A. Khan, M. F. Siddiqui and T. A. Khan, Ultrasonic-assisted synthesis of polyacrylamide/bentonite hydrogel nanocomposite for the sequestration of lead and cadmium from aqueous phase: equilibrium, kinetics and thermodynamic studies, Ultrason. Sonochem., 2020, 60, 104761 CrossRef CAS PubMed.
  103. S. I. Siddiqui and S. A. Chaudhry, Nigella sativa plant-based nanocomposite-MnFe2O4/BC: an antibacterial material for water purification, J. Clean. Prod., 2018, 200, 996–1008 CrossRef CAS.
  104. M. A. Ibrahim, M. Z. Jaafar, M. A. M. Yusof, C. A. Shye and A. K. Idris, Influence of size and surface charge on the adsorption behaviour of silicon dioxide nanoparticles on sand particles, Colloids Surf., A, 2023, 674, 131943 CrossRef CAS.
  105. T. Wollandt, S. Mangel, J. Kussmann, C. C. Leon, A. Scavuzzo, C. Ochsenfeld, K. Kern and S. J. Jung, Charging and Electric Field Effects on Hydrogen Molecules Physisorbed on Graphene, J. Phys. Chem. C, 2023, 127, 4326–4333 CrossRef CAS.
  106. E. Allahkarami, A. D. Monfared, L. F. O. Silva and G. L. Dotto, Toward a mechanistic understanding of adsorption behavior of phenol onto a novel activated carbon composite, Sci. Rep., 2023, 13, 1–16 CrossRef PubMed.
  107. O. Sahu and N. Singh, Significance of bioadsorption process on textile industry wastewater, The Impact and Prospects of Green Chemistry for Textile Technology, 2018, pp. 367–416 Search PubMed.
  108. S. Azizian and S. Eris, Adsorption isotherms and kinetics, Interface Sci. Technol., 2021, 33, 445–509 Search PubMed.
  109. N. Ayawei, A. N. Ebelegi and D. Wankasi, Modelling and Interpretation of Adsorption Isotherms, J. Chem., 2017, 1, 3039817 Search PubMed.
  110. P. Pourhakkak, A. Taghizadeh, M. Taghizadeh, M. Ghaedi and S. Haghdoust, Fundamentals of adsorption technology, Interface Sci. Technol., 2021, 33, 1–70 Search PubMed.
  111. K. H. Chu, Revisiting the Temkin Isotherm: Dimensional Inconsistency and Approximate Forms, Ind. Eng. Chem. Res., 2021, 60, 13140–13147 CrossRef CAS.
  112. Y. Liu, Y. Xiong, P. Xu, Y. Pang and C. Du, Enhancement of Pb (II) adsorption by boron doped ordered mesoporous carbon: Isotherm and kinetics modeling, Sci. Total Environ., 2020, 708, 134918 CrossRef CAS PubMed.
  113. N. Abbasi, S. A. Khan and T. A. Khan, Statistically optimised sequestration of mefenamic acid from polluted water by acacia gum phthalate/pectin hydrogel: a novel multifunctional adsorbent material synthesised via microwave-assisted process, Chem. Eng. J., 2023, 466, 143296 CrossRef CAS.
  114. H. K. Agbovi and L. D. Wilson, Adsorption processes in biopolymer systems: fundamentals to practical applications, Natural Polymers-Based Green Adsorbents for Water Treatment, 2021, pp. 1–51 Search PubMed.
  115. Y. Liu, New insights into pseudo-second-order kinetic equation for adsorption, Colloids Surf., A, 2008, 320, 275–278 CrossRef CAS.
  116. S. A. Chaudhry, Z. Zaidi and S. I. Siddiqui, Isotherm, kinetic and thermodynamics of arsenic adsorption onto Iron-Zirconium Binary Oxide-Coated Sand (IZBOCS): modelling and process optimization, J. Mol. Liq., 2017, 229, 230–240 CrossRef CAS.
  117. J. Wang and X. Guo, Adsorption kinetic models: physical meanings, applications, and solving methods, J. Hazard. Mater., 2020, 390, 122156 CrossRef CAS PubMed.
  118. S. Madan, R. Shaw, S. Tiwari and S. K. Tiwari, Adsorption dynamics of Congo red dye removal using ZnO functionalized high silica zeolitic particles, Appl. Surf. Sci., 2019, 487, 907–917 CrossRef CAS.
  119. F. Bessaha, N. Mahrez, K. Marouf-Khelifa, A. Çoruh and A. Khelifa, Removal of Congo red by thermally and chemically modified halloysite: equilibrium, FTIR spectroscopy, and mechanism studies, Int. J. Environ. Sci. Technol., 2019, 16, 4253–4260 CrossRef CAS.
  120. Z. Zaidi, A. Manchanda, A. Sharma, Shehnaz, A. Choudhry, M. Sajid, S. A. Khan, A. Khan and S. A. Chaudhry, Adsorptive removal of Methylene blue using fruit waste activated carbon and its binary metal oxide nanocomposite, Chem. Eng. J. Adv., 2023, 16, 100571 CrossRef CAS.
  121. S. Sircar, Adsorbate mass transfer into porous adsorbents – a practical viewpoint, Sep. Purif. Technol., 2018, 192, 383–400 CrossRef CAS.
  122. W. J. Weber Jr and J. C. Morris, Kinetics of Adsorption on Carbon from Solution, J. Sanit. Eng. Div., 1963, 89, 31–59 CrossRef.
  123. A. W. M. Ip, J. P. Barford and G. McKay, A comparative study on the kinetics and mechanisms of removal of Reactive Black 5 by adsorption onto activated carbons and bone char, Chem. Eng. J., 2010, 157, 434–442 CrossRef CAS.
  124. L. Ai, M. Li and L. Li, Adsorption of Methylene Blue from Aqueous Solution with Activated Carbon/Cobalt Ferrite/Alginate Composite Beads: Kinetics, Isotherms, and Thermodynamics, J. Chem. Eng. Data, 2011, 56, 3475–3483 CrossRef CAS.
  125. M. Afkari, S. M. Masoudpanah, M. Hasheminiasari and S. Alamolhoda, Effects of iron oxide contents on photocatalytic performance of nanocomposites based on g-C3N4, Sci. Rep., 2023, 13, 1–11 CrossRef PubMed.
  126. C. Li, S. He, H. Mo, X. Xu, P. Yang and M. Liu, Recent developments of iron oxide-based photocatalysts in water treatment technology: a review, Environ. Sci., 2025, 11, 1369–1385 CAS.
  127. M. Danish, M. S. Athar, I. Ahmad, M. Z. A. Warshagha, Z. Rasool and M. Muneer, Highly efficient and stable Fe2O3/g-C3N4/GO nanocomposite with Z-scheme electron transfer pathway: Role of photocatalytic activity and adsorption isotherm of organic pollutants in wastewater, Appl. Surf. Sci., 2022, 604, 154604 CrossRef CAS.
  128. S. Demirci, M. Yurddaskal, T. Dikici and C. Sarıoğlu, Fabrication and characterization of novel iodine doped hollow and mesoporous hematite (Fe2O3) particles derived from sol-gel method and their photocatalytic performances, J. Hazard. Mater., 2018, 345, 27–37 CrossRef CAS PubMed.
  129. C. N. C. Hitam and A. A. Jalil, A review on exploration of Fe2O3 photocatalyst towards degradation of dyes and organic contaminants, J. Environ. Manage., 2020, 258, 110050 CrossRef CAS PubMed.
  130. A. Halfadji, A. Chougui, R. Djeradi, F. Z. Ouabad, H. Aoudia and S. Rajendrachari, TiO2-Decorated by Nano-γ-Fe2O3 as a Catalyst for Efficient Photocatalytic Degradation of Orange G Dye under Eco-friendly White LED Irradiation, ACS Omega, 2023, 8, 39907–39916 CrossRef CAS PubMed.
  131. E. D'Souza, A. B. Fulke, N. Mulani, A. Ram, M. Asodekar, N. Narkhede and S. N. Gajbhiye, Decolorization of Congo red mediated by marine Alcaligenes species isolated from Indian West coast sediments, Environ. Earth Sci., 2017, 76, 1–13 CrossRef.
  132. B. S. Goud, H. L. Cha, G. Koyyada and J. H. Kim, Augmented Biodegradation of Textile Azo Dye Effluents by Plant Endophytes: A Sustainable, Eco-Friendly Alternative, Curr. Microbiol., 2020, 77, 3240–3255 CrossRef CAS PubMed.
  133. M. A. Sayed, A. Mohamed, S. A. Ahmed, A. M. El-Sherbeeny, W. Al Zoubi and M. R. Abukhadra, Enhanced photocatalytic degradation of Congo red dye into safe end-products over ZnO@polyaniline/coal composite as low cost catalyst under visible light: pathways and ecotoxicity, J. Photochem. Photobiol., A, 2024, 456, 115843 CrossRef CAS.
  134. M. Thomas, G. A. Naikoo, M. U. D. Sheikh, M. Bano and F. Khan, Effective photocatalytic degradation of Congo red dye using alginate/carboxymethyl cellulose/TiO2 nanocomposite hydrogel under direct sunlight irradiation, J. Photochem. Photobiol., A, 2016, 327, 33–43 CrossRef CAS.
  135. B. Li, X. Zhao, Y. Huang, X. Lu, H. Jia and M. Li, Dual Z-scheme BiOI/Bi2S3/MgIn2S4 composite photocatalyst for effective photocatalytic degradation of Congo red, Environ. Sci. Pollut. Res. Int., 2023, 30, 122537–122549 CrossRef CAS PubMed.
  136. G. He, Y. Li, W. Ji, Y. Hou, T. Zhu, Y. Chen, L. Yin and R. Liu, Highly efficient removal of Congo red from wastewater using biogenic vaterite: performance and mechanistic insights, Surf. Interfaces, 2025, 62, 106270 CrossRef CAS.
  137. C. Fei, C. Chen, H. Zhang, A. Liu, D. Ju, X. Yang, F. Wang, M. Xia and P. Huang, The effective removal of Congo red dye via chitosan-layered double hydroxide composite and the deep insight into mechanism, J. Ind. Eng. Chem., 2024, 135, 143–153 CrossRef CAS.
  138. H. Kim, C. Park, N. Choi and K. Cho, Congo red dye degradation using Fe-containing mineral as a reactive material derived from waste foundry dust, Environ. Sci. Pollut. Res., 2024, 31, 28443–28453 CrossRef CAS PubMed.
  139. L. G. Devi, S. G. Kumar, K. M. Reddy and C. Munikrishnappa, Effect of various inorganic anions on the degradation of Congo Red, a di azo dye, by the photo-assisted fenton process using zero-valent metallic iron as a catalyst, Desalination Water Treat., 2009, 4, 294–305 CrossRef CAS.
  140. J. O. Ighalo, F. O. Omoarukhe, V. E. Ojukwu, K. O. Iwuozor and C. A. Igwegbe, Cost of adsorbent preparation and usage in wastewater treatment: a review, Cleaner Chem. Eng., 2022, 3, 100042 CrossRef.
  141. Z. Sheerazi, I. Ahmad, S. A. Chaudhry, A. H. Sheikh and T. A. Khan, Green synthesis of a nicotinamide-functionalized cobalt ferrite nano-adsorbent for aqueous phase removal of some virulent dyes: evaluation of adsorption isotherms, kinetics, and mechanisms, New J. Chem., 2025, 49, 14590–14604 RSC.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.