Hydrothermal synthesis of AgInS2@biochar nanocomposites for the photocatalysis and electrochemical sensing of glufosinate herbicides

Firdous Ahmad Ganaiea, Irfan Nazira, Aaliya Qureashia, Zia ul Haqa, Kaniz Fatimaa, Arshid Bashirb, Altaf Hussain Pandith*a and Mohsin Ahmad Bhat*a
aLaboratory of Nanoscience and Quantum Computations, Department of Chemistry, University of Kashmir, Hazratbal, Srinagar-190006, J&K, India. E-mail: altafpandit23@gmail.com; mohsin@kashmiruniversity.ac.in; Fax: +91-194-2414049; Tel: +91-194-2424900 Tel: +91-7006429021 Tel: +91 9419033125
bDepartment of Chemistry, Degree College, Beerwah, Budgam, J&K, India

Received 24th May 2025 , Accepted 24th July 2025

First published on 25th July 2025


Abstract

A straightforward hydrothermal synthetic technique was used for the fabrication of a ternary chalcopyrite stacked biochar with great photocatalytic and electrocatalytic performance toward the photodegradation of malachite green (MG) dye and electrochemical sensing of glufosinate herbicides in contaminated water. The physicochemical characterization of the AgInS2@biochar nanocomposite analyzed using XRD, FTIR, TGA, SEM, TEM, BET, EIS, XPS, PL, and DR-UV spectroscopic techniques suggests that it is an electronically semiconducting composite matrix with an optical band gap of around 1.49 eV. Morphological studies demonstrated the formation of interlaced spherical nanospheres of AgInS2 nanoparticles with sufficiently exposed surfaces for effective surface activities. These include the photodegradation of the malachite green (MG) dye and electrochemical sensing of the glufosinate herbicide. Photodegradation studies reveal that AgInS2@biochar exhibits outstanding photocatalytic performance toward the photodegradation of malachite green (MG) from contaminated water, achieving a high removal efficiency of 97.4% within 70 min due to its high electron transfer efficiency. The degradation followed pseudo-first-order kinetics, with a high K1 value of ca. 0.0433 min−1. In addition, the presented voltammetric studies establish that AgInS2@biochar is a favourable electrode material for elusive electrochemical sensing of the toxic herbicide glufosinate in aqueous solutions with a low detection limit of ca. 0.026 μM.


1. Introduction

Globally, we are facing a scarcity of drinking water. This is largely due to the presence of various contaminants that are directly or indirectly discharged into water bodies.1–3 These contaminants include potentially hazardous dyes, herbicides, heavy metal ions, and microplastics.4,5 Particularly, potentially hazardous dyes, such as malachite green, methylene blue, and organophosphate herbicides, are the major contributors to water contaminants.6 Therefore, it is essential to treat wastewater before its discharge into the environment.7,8 There is an urgent need for cost-effective and highly efficient materials for water treatment applications. A variety of semiconductors, including ZnO, CuO, CdS, ZnS, TiO2, SnO2, Ag3PO4, silver nanoparticles (Ag-NPs), and titanium oxide nanoparticles (TiO2-NPs), have been utilized as photocatalysts for effective water purification. Despite their widespread use, some of these materials exhibit limited activity under visible light and are susceptible to photo-corrosion.9–12 Hence, there is a need for innovative approaches that can address both the detection and removal aspects of contamination. Researchers around the world are continuously working to develop and modify new materials that enhance photocatalysis and electrochemical sensing of various contaminants.13,14 Hybrid heterostructures can be used in electrochemical sensing applications utilizing the complementary properties of various materials. In this context, ternary semiconductor materials, such as TbCrO3, CuInS2, CuInSe2, and AgInS2, have gained significant attention owing to their low toxicity.15,16 Among these, AgInS2 stands out due to its remarkable properties, including a narrow emission band, broad excitation band, long fluorescence lifetime, high fluorescence emission, and excellent photostability.17–19 It is a non-toxic and highly promising material, particularly effective as a solar light absorber for degrading organic pollutants. The unequal bonds between AgS and InS are beneficial for the formation of an internal electric field. Therefore, charges can be separated rapidly under illumination, leading to an improvement in the photocatalytic activity.20–22 Despite its various advantages, AgInS2 faces limitations, such as photocorrosion, charge recombination, tendency to agglomerate during reactions, and limited adsorption capacity, due to its relatively small specific surface area. To address these challenges, AgInS2 has been combined with carbon materials, such as biochar, to produce a heterostructure, which can expand the light absorption range and promote charge separation to improve its effectiveness in treating toxic contaminants.23–26 These properties are considered highly effective in reducing charge recombination and enhancing photocatalytic activity.27–29

Biochar has evolved as an efficient and functional support material for the decoration of nanocatalysts. Its ease of preparation, chemical stability, versatile functional groups, electrical conductivity, large surface area, cost-effectiveness and unique properties make it suitable for various applications.30–32 We fabricated an AgInS2@biochar nanocomposite via a novel hydrothermal route utilizing cattail (Typha plant) biomass as a green source of biochar. The nanocomposite offers heterogeneous interfaces, flexible band gaps, and superior charge separation ability. The resulting synergy enhances sensitivity, selectivity, versatility, and overall performance, making these structures pivotal for the advancement of photocatalytic and electrochemical sensing technology.33–35 Various analytical techniques have been utilized for the detection of contaminants, including gas chromatography (GC),36 capillary electrophoresis,37 mass spectrometry,38 UV-visible spectrometry,39 and high-performance liquid chromatography (HPLC).40 Among these, the electrochemical method stands out owing to its operational simplicity, high sensitivity, and selectivity, making it particularly effective for real-time monitoring of environmental and food samples.41–43

In this study, we develop a novel strategy for the design and synthesis of AgInS2 and its composite (AgInS2@biochar) using a two-step hydrothermal method. The biomass collected from the cattail typha plant (Typha latifolia L.) was pyrolyzed to obtain a biochar material that served as a sustainable carbon source. Both AgInS2 and AgInS2@biochar materials were employed for catalytic and sensing applications. Although pristine AgInS2 demonstrated limited photocatalytic activity and a weak electrochemical response (CV and DPV), the AgInS2@biochar composite exhibited significantly enhanced electrochemical sensitivity toward the detection of the toxic herbicide in 0.1 M KCl with glufosinate concentrations ranging from 1 to 10 μM, achieving a detection limit of approximately 0.026 μM-well within the limits recommended by the WHO. The AgInS2@biochar sensor also demonstrated good stability, reusability, and improved detection limits compared to previously reported materials. Furthermore, the effective photocatalysis of malachite green (MG) dye was carried out with high photocatalytic activity. To the best of our knowledge, this is the first report on the electrochemical sensing properties of AgInS2@biochar for herbicide detection.

2. Materials and methods

2.1 Materials

Cattail (Typha) plants were collected from the banks of Dal Lake, Srinagar; their chemical name is Typha latifolia L. with 95% purity, while silver nitrate (AgNO3·5H2O, 97.5%) was purchased from High Purity Laboratory Chemicals Pvt. Ltd. Indium chloride (InCl3), thioacetamide (98.0%), and malachite green dye were purchased from MERCK. Glufosinate herbicide was obtained from Cheminova Pvt. Ltd. Double-distilled water was used to prepare the solutions in this study.

2.2 Preparation of biochar

The locally sourced cattail was initially desiccated at a temperature of 30 °C using a vacuum oven. Subsequently, the air-dried cattail underwent crumpling and filtration. Following this, the dehydrated cattail material was placed in a crucible and subjected to slow pyrolysis in a tubular furnace at a heating rate of 10 °C per minute for 2 h. This process resulted in the production of a dark black biochar material, which was then collected and stored for future experimental use.

2.3 Preparation of AgInS2 nanocomposite material

AgInS2 nanospheres were synthesized using a simple hydrothermal method. Specifically, 0.3 mmol of AgNO3, 0.5 mmol of InCl3·4H2O, and an excess of thioacetamide were dissolved in 20 mL of deionized water for approximately 30 min. This solution was then transferred to a 50 mL Teflon-lined stainless-steel autoclave and kept at 120 °C for 8 h in a hot air oven. The resulting brownish-black colored powder was collected, washed, and dried at 50 °C.

2.4 Preparation of the AgInS2@biochar composite

The AgInS2@biochar nanocomposite was prepared following a hydrothermal method. Initially, 0.20 g of Cattail biochar was dispersed in 20 mL of distilled water under ultra-sonication for 1 h. To this biochar dispersion, 0.3 mmol silver nitrate (AgNO3), 0.5 mmol indium tri chloride (InCl3), and 2 mmol thioacetamide were added. The whole mixture was then stirred for 1 h and transferred to a hydrothermal reaction vessel maintained at 120 °C for 8 h. The precipitate was filtered and repeatedly rinsed with distilled water. The precipitate was then dried at 50 °C for 2 h to produce a solid black powder of AgInS2@biochar (Scheme 1).
image file: d5nj02188j-s1.tif
Scheme 1 Synthetic scheme of AgInS2@biochar.

2.5 Catalyst characterisation

An X-ray diffractometer (XRD), Ultima-IV, Rigaku Corporation, Tokyo, Japan, was used to probe crystallinity. The morphology was examined with an analytical scanning electron microscope (Zeiss Ultraplus-4095) and a transmission electron microscope (HR-TEM) using a JEOL JEM 2100F (USA). Diffuse reflectance spectra (DRS-UV) were recorded with a Shimadzu 2600 spectrometer (Japan). The surface functionalization was analyzed with FTIR spectroscopy using a PerkinElmer Spectrum-100 FTIR. Photoluminescence (PL) spectra were recorded at room temperature with a Hi-tech F-4600 fluorescence spectrophotometer (λmax = 440 nm). The X-ray photoelectron spectra of the samples were analyzed using a PHI 5000 Versa Probe III XPS instrument with monochromatized Al Kα (E = 1486.7 eV).

2.6 Photocatalytic activity test

The photocatalytic properties of the AgInS2@biochar composite were evaluated by monitoring the degradation of malachite green (MG) using a flow batch mode method. A 300-W UVB lamp (Philips) was utilized in a photoreactor setup. The experiment involved adding 1 mg mL−1 of the nanocomposite to 100 mL of a 10 ppm MG solution. This mixture was stirred in the dark for 1 h to reach adsorption–desorption equilibrium before irradiation. During the photocatalytic reaction, 2.5 mL aliquots were taken every 10 min and centrifuged to remove photocatalyst particles. The absorbance at 614 nm was measured using a double-beam UV-visible spectrophotometer (model: LI-2802). The degradation efficiency of MG was calculated using the following equation:
 
image file: d5nj02188j-t1.tif(1)
where the absorbances C0 and C are the concentrations of malachite green (MG) before and after exposure to visible light, respectively. The experimental data were analyzed to determine the kinetics of MG adsorption and degradation using the pseudo-first-order kinetic (PFOK) model. This model is represented by the equation typically used to describe the rate of reaction in photocatalytic processes.
 
image file: d5nj02188j-t2.tif(2)
where t is the given time (min), [MG0] is the initial concentration of MG (mg L−1), and [MG]t is the concentration of MG at another given time (mg L−1).

2.7 Electrochemical studies

Electrochemical studies were conducted using a bio-logic-sas potentiostat (SP-150) for cyclic voltammetry (CV) and differential pulse voltammetry (DPV) in a three-electrode configuration. The working electrode was a 3-mm diameter glassy carbon electrode (GCE) that had been modified with the synthesised material. The counter and reference electrodes were platinum wire and Ag/AgCl (3 M KCl), respectively. The studies used 0.1 M KNO3 solution as the supporting electrolyte, with a potential window ranging from −1 V to +1 V and a scan rate of 5 mV s−1. The electrolyte solution was purged with high-purity argon/nitrogen gas at atmospheric pressure (∼1 atm) for 20 minutes to eliminate dissolved oxygen and ensure an inert atmosphere. In addition, the glassy carbon electrode (GCE) was polished with alumina slurry (0.05 μm) on a polishing cloth. It was then thoroughly rinsed with distilled water and ethanol before drying at an ambient temperature. To modify the electrode, a stable suspension of the AgInS2@biochar nanocomposite was created by dispersing 2 mg of the material in 1 mL of ethanol and 10 μL of 0.5% glutaraldehyde solution, followed by sonication for 30 minutes to ensure uniform dispersion. To create a homogenous and adherent nanocomposite film, 5 μL of the well-dispersed solution was drop-cast onto a clean GCE surface and dried at room temperature. A modified electrode (AgInS2@biochar/GCE) was employed for electrochemical sensing of the glufosinate herbicide, and DPV was employed to evaluate this sensitivity. The limit of detection (LOD) was calculated using the following standard equation:
 
image file: d5nj02188j-t3.tif(3)
where sigma is the standard deviation in experimental current values and S represents the slope in the calibration plot.

2.8 Computational analysis

Density functional theory (DFT) was employed in the computational investigations along with the Gaussian 09 arrangement of codes. B3LYP and LanL2DZ were applied as the practical and basis sets, respectively. Geometry optimization was carried out using Becke's three-parameter exchange functional combined with the Lee–Yang–Parr correlation function. The LanLDZ/6-311++g(d,p) basis sets were used for AgInS2 and non-metal atoms, such as carbon. Moreover, the HUMO, LUMO and HUMO–LUMO energy gaps were assessed for AgInS2 and AgInS2@biochar.

3. Results and discussion

3.1 Characterization of the material

X-ray diffraction studies were conducted to analyze the phase and crystallinity of the as-prepared materials. The PXRD diffraction patterns for biochar, AgInS2, and AgInS2@biochar are presented in (Fig. 1a). The biochar samples display sharp diffraction peaks at approximately 2θ = 22.3° and 28.5°, corresponding to the (100) and (002) planes of the carbon structure, respectively. In AgInS2, diffraction peaks at 24.3°, 28.2°, 31.2°, 34.0°, 40.2°, 49.8°, and 59.7° and 71° were observed. These results correspond to an orthorhombic structure (JCPDS No. 25–1328).44,45 The AgInS2@biochar composite exhibits peaks corresponding to both pristine biochar and AgInS2. Notably, the peaks significantly deviate from their standard positions, and the presence of higher-intensity peaks indicates an increased number of atoms occupying the same lattice planes. This deviation and intensity suggest the successful formation of the AgInS2@biochar nanocomposite through biochar impregnation.
image file: d5nj02188j-f1.tif
Fig. 1 (a) XRD spectra of biochar, AgInS2, and AgInS2@biochar, (b) FT-IR spectra of biochar, AgInS2, and AgInS2@biochar, (c) TGA of biochar, AgInS2, and AgInS2@biochar, and (d) BET N2 adsorption–desorption isotherms of ZnIn2Se4@biochar.

In addition, the average particle size in AgInS2 and AgInS2@biochar was calculated using the Debye–Scherrer equation:

d = /β[thin space (1/6-em)]cos[thin space (1/6-em)]θ,
where d = grain size, λ = 1.54 Å (wavelength of the X-ray radiation), K = 0.89, β = peak width (in radians) at half-maximum height, and 2θ = Bragg's angle. For orthorhombic phase AgInS2, the average grain size in AgInS2 and AgInS2@biochar were calculated to be ca. 27.3 and 38.2 nm, respectively.

FTIR analysis of the crafted samples was carried out to determine the functional groups present over the surface of the functional groups in biochar, AgInS2, and AgInS2@biochar nanocomposite (Fig. 1b). In all three samples, the peaks located around 3480 cm−1 were attributed to the strong stretching vibrations of O–H. The peak at 2840 cm−1 corresponds to a weak C–H stretching vibration band. In biochar and AgInS2@biochar, a peak at around 1630 cm−1 is attributed to the presence of C[double bond, length as m-dash]O in the sample. The peaks at 1520 cm−1 and 1649 cm−1 are due to the bending vibrations of –CH3, the characteristic of the carbon material. The stretching bands at approximately 564 cm−1, 982 cm−1, and 1205 cm−1 could be due to In–S or Ag–S bonds.44 Moreover, the bands of AgInS2 are present in the composite material (AgInS2@biochar), indicating that AgInS2 has been successfully incorporated into the biochar Matrix.

TGA analysis was carried out to determine the thermal stability of the synthesized materials (Fig. 1c). The biochar sample undergoes a two-step weight loss, with a 15% reduction between 100 °C and 150 °C, and a further weight loss starting from 590 °C to 700 °C results in an overall 23% weight reduction due to the decomposition of cellulosic, hemicellulose, and lignin compounds, after which the weight stabilizes.46 In contrast, AgInS2 starts losing weight around 150 °C, continuing up to 500 °C, which results in an 80% weight loss primarily due to the evaporation of adsorbed moisture. In AgInS2@biochar the removal of adsorbed water is observed upto 100 °C. Another weight loss is observed upto 260 °C due to loss of adsorbed water and decomposition of organic matter in the composite. Thereafter the weight loss continues upto 540 °C due to decomposition of organic matter. Compared to pristine biochar, the composite exhibit delayed and reduced weight loss. This indicate the superior thermal stability of the composite material. The AgInS2@biochar composite starts losing weight around 100 °C, continuing up to 280 °C, resulting in an 27% weight loss and further, continuing occurs up to 540 °C results 38% weight loss exhibits delayed and reduced weight loss, indicating superior thermal stability compared to AgInS2, attributed to the expulsion of surface-bound water and decomposition of components other than carbon.

The surface area of the biochar and AgInS2@biochar composite was measured using the multipoint Brunauer–Emmett–Teller (BET) method based on nitrogen adsorption–desorption isotherms. The isotherms displayed in (Fig. 1d) show a type-IV pattern with a hysteresis loop at a relative pressure (P/P0) from 0.0 to 1.2, indicating a mesoporous structure. The BET analysis revealed a specific surface area of pristine biochar is 17.3 m2 g−1 (ref. 47) and that of the composite is 28.1 m2 g−1, confirming its advanced porous structure. This enhanced surface area provides a larger contact region, which is advantageous for photocatalytic and electrochemical sensing applications.

Scanning electron microscopy (SEM) was used to examine the surface morphology of biochar, AgInS2, and the AgInS2@biochar composite, as shown in (Fig. 2). The biochar exhibits interconnected bar-like structures resembling hollow tubes with a reticular surface (Fig. 2a and b). This surface structure of biochar is desirable for AgInS2 to reunite into a spherical-like morphology, quick mass transport, and the growth of AgInS2 interlaced spherical nanospheres25 (Fig. 2c). In the AgInS2@biochar composite (Fig. 2d), AgInS2 appears as nanosphere units and particulate deposits on the biochar surface, suggesting interface interactions between the components. The heterostructure shows significant aggregation akin to the pristine composite. Transmission electron microscopy (TEM) analysis (Fig. 2e and f) reveals that the AgInS2 particles were decorated on the interface of the composite to design AgInS2@biochar hybrids without any destruction of the biochar or AgInS2 nanospherical structures. We observe the successful stacking of AgInS2 on the highly reticular surface of biochar, confirming the efficacious production of the composite material.


image file: d5nj02188j-f2.tif
Fig. 2 (a) and (b) SEM images of biochar, (c) AgInS2, and (d) composite AgInS2@biochar and (e) and (f) TEM images of AgInS2@biochar.

The EDS analysis was performed to assess the presence and distribution of elements in the AgInS2@biochar composite. The results confirm a uniform distribution of all elements, specifically carbon (C), oxygen (O), silver (Ag), indium (In), and sulfur (S). These significant findings are illustrated in (Fig. 3).


image file: d5nj02188j-f3.tif
Fig. 3 (a) EDS spectra of AgInS2@biochar (b–f) EDS elemental mapping of AgInS2@biochar.

3.2 Optical studies

UV-Vis diffuse reflectance spectroscopy was used to analyse the solid samples using a Shimadzu 2600 spectrometer (Japan), equipped with a 240-52454-01 integration sphere accessory. The solid material was ground in a pan grinder to a 100-mesh size. Then, this material was analyzed after being compressed carefully into the spectrometer cell. The DRUV-VIS spectra were converted to the Kubelka–Munk remission function defined by f(KM) = (1 − R)2/2R = k/s, where R is the reflectance, k is the absorption coefficient, and s is the scattering coefficient. Moreover, UV-vis diffuse reflectance spectroscopy was employed to analyze the optical properties of the synthesized materials, confirming the semiconductor characteristics of both AgInS2 and AgInS2@biochar composite, as depicted in (Fig. 4). AgInS2 exhibited extended absorption in the visible range around 540 nm, while the AgInS2@biochar composite showed a redshift of the absorption edge to approximately 685 nm. Band gaps were determined from Tauc plots of (αhν)2 versus hν (Fig. 4b), revealing a value of 1.89 eV for AgInS2 and 1.49 eV for the AgInS2@biochar composite. The incorporation of biochar results in a lower band gap energy compared to pristine AgInS2, which is attributed to enhanced charge transfer.
image file: d5nj02188j-f4.tif
Fig. 4 (a) DR-UV spectra of biochar, AgInS2, and AgInS2@biochar. (b) Tauc plots of the AgInS2@biochar. (c) The PL spectra of AgInS2 and AgInS2@biochar.

Photoluminescence (PL) analysis assessed the recombination rate of solar light-generated electron–hole pairs. (Fig. 4c) shows the PL spectra for AgInS2 and AgInS2@biochar, both excited at a wavelength of 670 nm, with an emission peak around 685 nm due to edge-free excitation luminescence. The PL intensity of the AgInS2@biochar composite is notably lower than that of AgInS2 alone, indicating that forming the composite effectively reduces the recombination of photo-induced carriers and enhances electron–hole pair separation. Biochar acts as an electron trap, improving electron trapping and resulting in a smaller PL peak for the composite. This suggests that AgInS2@biochar composites can extend the lifespan of holes (h+) and activate hydroxyl groups to form ˙OH radicals on the biochar surface, aiding in the degradation of water pollutants.

3.3 XPS studies

X-ray photoelectron spectroscopy (XPS) analysis was performed to examine the chemical state of the synthesized AgInS2@biochar composite. The survey spectrum revealed the presence of C 1s, O 1s, Ag 3d, In 3d, and S 2p signals, with no impurities observed (Fig. 5a). The weak presence of oxygen over the surface, as reflected by the XPS spectra, can be attributed to the held oxygen-containing bundles on the surface of the biochar. The C 1s spectra of the biochar, depicting the two different peaks at 284.6 and 286.1 eV, are attributed to C–C and C–O–C, respectively (Fig. 5b). The binding energies of the O 1s spectrum in (Fig. 5c) showed that the deconvoluted peaks at 531.5 and 529.2 eV were attributed to the O atoms ascribed to C–O and C[double bond, length as m-dash]O bonds, respectively. High-resolution spectra were also taken for the Ag 3d region (Fig. 5d), and the Ag 3d core splits into 3d5/2 at (368.1 eV) and 3d3/2 at (372.5 eV) peaks correspondingly, with a spin–orbit splitting of 4.4 eV.25 In the high resolution In 3d spectra, the binding energies of In 3d5/2 and In 3d3/2 were 444.5 eV and 452.1 eV, respectively (Fig. 5e). In addition, the S 2p spectra correspond to two peaks of S 2p3/2 and S 2p1/2 peaks at 161.3 and 162.2 eV, that are assigned to S coordinated with Ag and In in AgInS2, respectively (Fig. 5f). The binding energies of Ag 3d, In 3d and S 2p in AgInS2 are consistent with those reported in the literature.25,26 The current XPS analysis confirmed the presence of all elements in the composite, demonstrating the fruitful spreading of AgInS2 on the surface of the biochar.
image file: d5nj02188j-f5.tif
Fig. 5 (a) Complete XPS range, (b) C 1s XPS, (c) O 1s XPS, (d) Ag XPS, (e) In XPS, and (f) S XPS.

4. Zeta potential and impedance studies

The zeta potential (Fig. 6a) of AgInS2@biochar nanocomposite becomes more negative at basic pH due to the deprotonation of surface functional groups and the increased presence of negatively charged species. This results in a higher negative charge density on the nanocomposite surface, thereby increasing the negative zeta potential. The more negative zeta potential enhances the electrostatic repulsion between the nanocomposite particles, improving the stability of the colloidal suspension and influencing the sensitivity of the sensor and photocatalytic efficiency. For electrochemical sensing of glufosinate, this negative charge can facilitate better interaction with positively charged analytes.
image file: d5nj02188j-f6.tif
Fig. 6 (a) Apparent zeta potential. (b) EIS Nyquist plots of AgIns2 and AgInS2@biochar.

Moreover, impedance studies reveal that the AgInS2-coated biochar exhibits superior charge transport abilities compared to pristine AgInS2. (Fig. 6b) illustrates sample Nyquist plots corresponding to the EIS data recorded over AgInS2 and AgInS2@biochar. The depicted Nyquist plots seem typical of those expected from a Rendell's circuit, wherein the radius of the semicircle noted in the high-frequency region is taken as a measure of the resistance to charge transfer. The Nyquist plot shows a remarkably smaller semicircle radius for the AgInS2@biochar composite compared to AgInS2, indicating superior kinetics of heterogeneous electron transfer in the composite. The data were fitted to an electrochemical circuit model to estimate charge transfer resistance (Rct). For AgInS2@biochar, Rct is 45.2 ohms, while for AgInS2 on a glassy carbon electrode (GCE), it is 87.4 ohms, reflecting higher resistance and slower electron movement.48,49 The lower Rct for AgInS2@biochar suggests that biochar integration significantly enhances electron transfer, thereby improving the composite's photocatalytic and electrocatalytic performance.

5. Photocatalytic study of MG

AgInS2@biochar was explored for the photocatalytic degradation of malachite green (MG). First, 10 mg of the AgInS2@biochar nanocomposite was added to a 100 mL solution of 10 ppm of MG. Before irradiation, the catalyst and MG solution were stirred for 1 h to facilitate adsorption on h-AgInS2@biochar. Aliquots were periodically taken to monitor degradation via UV-vis spectrophotometry. The malachite green has an absorption peak at 614 nm, and the decreasing intensity of absorbance corresponding to MG specific in the depicted UV-visible spectra suggests the photocatalytic degradation of MG (Fig. 7a). The catalyst was found to degrade more than 97.4% of the MG within 70 min (Fig. 7b). The enhanced performance is attributed to the synergistic interaction between AgInS2 and biochar that boosts the visible-light-harvesting ability of the nanocomposite owing to the surface plasmon resonance phenomenon and the graphitic structure of the carbon that increases the interfacial charge separation, thus decreasing the recombination of electron–hole pairs, which in turn increases the generation of reactive oxygen species (ROS) in the photocatalytic degradation procedure. In (Table 1), we compared the photocatalytic results of AgInS2@biochar and other photocatalyst materials available in the literature. The results dictate the superior performance of AgInS2@biochar. Subsequently, it revealed great photocatalytic performance. The kinetics (Fig. 7c) follow a pseudo-first-order model (eqn (2)) with a rate constant of k = 0.0433 min−1.
image file: d5nj02188j-f7.tif
Fig. 7 (a) Photocatalytic degradation of MG. (b) Percentage degradation graph. (c) The Langmuir–Hinshelwood model of ln(C0/C) function with time. (d) Mechanism of dye degradation.
Table 1 Comparison table for photocatalytic degradation of MG dye using different photocatalysts
Materials Irradiation time (min) Degradation % of MG Ref.
Ag/ZnO/g-C3N4 60 91 50
ZnO/CNT 60 79 51
ZnO-La2CuO4 120 91 52
Zno@graphene oxide 120 97 53
CuO@GO 60 99.6 54
SnS (NPs) 75 98 55
Ag/ZnO polymeric nanofibers 60 93.5 56
CuO/ZnTe 97.2 80 57
WT-ZnO NPs 105 81.5 58
AgInS2@biochar 120 97.4 This work


5.1 Adsorption study

To endorse the influence of the adsorption of MG onto AgInS2@biochar on the total removal of MG, we performed the adsorption studies in a batch mode system at room temperature. Concisely, the 10 mg L−1 MG solution was agitated with AgInS2@biochar in the dark for a 70 min contact time. At the end of the adsorption period, the solution was centrifuged for 10 min at 5000 rpm. After centrifugation, small amounts of the liquid were taken and analysed using UV-vis absorption spectroscopy (Fig. 8a). The results specify a considerable decline in peak intensity. This suggests the dynamic role played by the process of adsorption in the overall mineralisation of MG. We report 24.8% MG removal by adsorption alone, as represented in (Fig. 8b). Hence, the catalyst material bids a sufficient surface for the effective adsorption of pollutants.
image file: d5nj02188j-f8.tif
Fig. 8 (a) Decrease in the absorption intensity with time in the absence of sunlight irradiation. (b) Removal percentage of MG by adsorption.

5.2 Mechanism of photocatalytic degradation of MG

In the photocatalytic mechanism, the light excitation of the photocatalyst material triggers the photo generation of radical species, eventually resulting in the degradation of the organic contaminants. In the present study, the photo irradiation of the AgInS2@biochar composite is expected to result in the generation of electron hole pairs and finally the radical species59 following eqn (4)–(8). In the photodegradation process of malachite green (MG), the primary mechanism involves the generation of electron–hole pairs through the photoexcitation of valence band electrons in AgInS2@biochar, which are then promoted to the conduction band. The energy levels of these bands facilitate the oxidation of water by the holes in the valence band, resulting in the production of hydroxyl radicals (˙OH). Concurrently, the photogenerated electrons in the conduction band possess sufficient reducing power to interact with surface oxygen, leading to the formation of superoxide radicals (O2˙). These radicals, O2˙ and ˙OH, collaboratively degrade MG into smaller fragments, eventually converting it into carbon dioxide (CO2) and water (H2O). (Fig. 7d) illustrates a plausible sequence of steps involved in the AgInS2@biochar-mediated photodegradation of malachite green (MG).
 
AgInS2@biochar + → AgInS2@biochar h+(VB) + e(CB) (4)
 
AgInS2@biochar h+(VB) + H2O → AgInS2@biochar + H+ + OH (5)
 
AgInS2@biochar − OHad + h(VB)+ → AgInS2@biochar + OH˙ad (6)
 
AgInS2@biochar (eCB) + O2 → AgInS2@biochar O2˙ (7)
 
MG + (OH˙, OH˙ad,O2˙−,or[thin space (1/6-em)]hVB+) → intermediates CO2/H2O) (8)

To further identify the degradation products of MG dye catalyzed by AgInS2@biochar, we propose a probable degradation mechanism of MG catalyzed by AgInS2@biochar. As shown in Scheme 2, the oxidation of MG at the tertiary carbon atom produces two main products: (4-(dimethylamino)phenyl) (phenyl)methanone and 4-(dimethylamino) phenol as well as N,N-dimethyl-4-((4-(methylamino)cyclohexa-2,5-dien-1-ylidene) methyl) aniline, via cleavage of the benzene group. Besides, N-demethylation reaction occurred for MG, creating the product of N-methyl-N-(4-((4-(methylamino)phenyl) (phenyl)methylene)cyclohexa-2,5-dien-1-ylidene)methanaminium, which may also undergo oxidation, and the oxidation products further undergo N-demethylation as well. Note that both oxidation and N-demethylation reactions were reported in the literature for MG.60,61


image file: d5nj02188j-s2.tif
Scheme 2 Proposed degradation mechanism for MG catalyzed by AgInS2@biochar based on the reported literature.

6. Electrochemical studies

To assess the efficacy of the synthesized materials as electrode components, detailed voltammetric analyses were performed using AgInS2- and AgInS2@biochar-modified glassy carbon electrodes (GCE) in an argon-saturated 0.1 M KNO3 electrolyte solution. (Fig. 9a) illustrates cyclic voltammograms (CVs) recorded at a potential scan rate of 20 mV s−1 at 298.15 K, comparing bare GCE with those modified by AgInS2 and AgInS2@biochar composites in 0.1 M KNO3. The CVs for the bare GCE show no faradaic signals across the potential window in the absence of glufosinate, while significant faradaic signals are evident in the CVs for the modified electrodes. Additionally, (Fig. 9b) presents CVs for the AgInS2@biochar-modified GCE in 0.1 M KNO3 with glufosinate recorded at various potential scan rates. An increase in scan rate was observed to increase the glufosinate-specific peak currents, suggesting diffusional control over the mass transfer of the herbicide. Interestingly, the CVs recorded over AgInS2@biochar in the presence of glufosinate exhibited a single cathodic peak in the forward scan and one anodic peak, with peak positions that precisely match the anodic peaks noted in the CV records for the toxic herbicide. This suggests the potential of AgInS2@biochar to facilitate sensitive and selective electrochemical sensing of glufosinate, the toxic herbicide. This was further confirmed by our detailed DPV investigations carried out on the AgInS2@biochar-modified GCE for differently concentrated solutions of glufosinate in 0.1 M KNO3. We conducted differential pulse voltammetric studies of AgInS2@biochar in various concentrations of glufosinate (1–10 μM). The oxidation current increases with the concentration of glufosinate in the solution (Fig. 9c). The calibration plot reveals a linear relationship between the current and glufosinate concentration, with a correlation coefficient (R2) of 0.995, indicating high sensitivity towards glufosinate herbicide (Fig. 9d). The limit of detection was determined to be 0.026 μM, demonstrating that AgInS2@biochar is effective for detecting glufosinate ions in solution. Both cyclic voltammetry and DPV confirm that AgInS2@biochar nanoparticles are suitable for sensing glufosinate. The detection capability of the AgInS2@biochar-modified glassy carbon electrode (GCE) for glufosinate is superior to that of other modified GCEs, as shown in Table 2.
image file: d5nj02188j-f9.tif
Fig. 9 (a) Cyclic voltammograms of biochar and AgInS2@biochar. (b) Modified GCE in 10 ppm solution of glufosinate ions at various scan rates. (c) Differential pulse voltammograms showing an increase in peak current with an increase in concentrations of glufosinate. (d) Limit of detection.
Table 2 Sensing of pesticides compared with our bio-based composite material
Materials Sensing of pesticides Detection limit (μM) Ref.
CuInS2 QDs Glufosinate 0.17 62
CuO/Fe2O3 nanozymes Glufosinate 28 63
Gold nano bipyramids Glyphosate 5.92 64
AuNPs with the enzyme urease Glyphosate 2.96 65
NiAl-LDH Glyphosate and glufosinate 1 66
AgInS2@biochar Glufosinate 0.026 This work


6.1 Possible interactions of hlufosinate with AgInS2@biochar

The interaction between glufosinate and AgInS2@biochar likely involves multiple mechanisms. Electrostatic attractions may play a significant role, as the charged groups on glufosinate can interact with the oppositely charged surface of AgInS2@biochar. Hydrogen bonding is also probable due to the presence of hydrogen bond donors and acceptors in both glufosinate and the biochar matrix. The porous structure of biochar facilitates diffusion, allowing glufosinate molecules to penetrate and interact with internal surfaces. Electrocatalysis could be involved, where the AgInS2 component of the biochar promotes electron transfer reactions, which enhances the interaction, as evidenced by the increase in the redox peak current, thereby facilitating the sensing of glufosinate. Additionally, complexation may occur, with glufosinate forming stable complexes with metal ions present in AgInS2@biochar. These combined mechanisms contribute to the effective binding and interaction of glufosinate with AgInS2@biochar (Scheme 3).
image file: d5nj02188j-s3.tif
Scheme 3 Probable mechanism of glufosinate with AgInS2@biochar.

6.2 Density functional theory analysis

DFT calculations were conducted utilising the Lee–Yang–Parr (B3LYP) functional, the LanL2DZ basis set for the AgInS2 atom, and the 6-311G(d,p) basis set for non-metal atoms to analyse the electronic characteristics of the materials. Utilising the Gaussian 03 arrangement of projects, we used the DFT method to determine their band hole energy values to focus on the effect of AgInS2 on biochar. The HUMO–LUMO band hole energies were determined to be 2.54 eV and 1.58 eV for AgInS2 and AgInS2@biochar, respectively. This significant reduction in the band gap during the formation of the composite designates increased electron mobility, as a lower HOMO–LUMO gap correlates to a reduced activation overpotential for electron transfer between the electrode and analytes. The computed band gap values and orbital distributions are shown in (Fig. 10). The band hole energy values were in close agreement with the exploratory qualities, as displayed in Table 3. These results show that the association of AgInS2 on the outer layer of biochar improved the optical response of the material. These outcomes indicate that the incorporation of AgInS2 on the surface of biochar enhanced photo absorption towards the visible region, which in turn enhances the charge separation of photo-excited electrons.
image file: d5nj02188j-f10.tif
Fig. 10 (a) AgInS2 and (b) AgInS2@biochar HUMO and LUMO energy gap calculated by DFT.
Table 3 Band gap energy values of biochar and the prepared composites
Name HUMO (eV) LUMO (eV) Band gap (eV)
AgInS2 −6.83 −4.29 2.54
AgInS2@biochar −5.956 −4.37 1.58


7. Selectivity and interference study

To assess the selectivity of the developed sensor towards glyphosate, cyclic voltammetry (CV) responses were recorded in the presence of glyphosate and other common interfering pesticides, including mancozeb, carbendazim, and tebuconazole (Fig. 11a). As shown in the CV plot, the sensor displayed a significantly distinct and well-defined redox response for glyphosate compared to the other analytes. The current response and peak shape for glyphosate were markedly different, indicating minimal interference from coexisting species. This suggests a high degree of selectivity of the sensor for glyphosate likely due to favorable binding interactions or specific affinity towards functional groups present in glyphosate. These results validate the applicability of the sensor in real sample analysis, even in complex matrices containing multiple agrochemical residues.
image file: d5nj02188j-f11.tif
Fig. 11 (a) Selectivity of the developed sensor towards glyphosate, mancozeb, carbendazim, and tebuconazole. (b) Repeatability and operational stability of the fabricated electrode were evaluated by measuring the current response over a period of seven consecutive days.

The repeatability and operational stability of the fabricated electrode were evaluated by measuring the current response for seven consecutive days. As shown in (Fig. 11b) in the bar graph, the current density remained nearly constant with only minor fluctuations, and all values were within a narrow error range (±0.002 mA cm−2). This consistent performance indicates the excellent repeatability and stability of the electrode under ambient storage conditions. The negligible variation in electrochemical response over time confirms that the sensor material does not degrade or lose activity, demonstrating its reliability for long-term applications in environmental sensing.

8. Conclusion

The novel bio-based hydrothermal synthesis of the AgInS2@biochar nanocomposite has demonstrated significant potential in both photocatalysis and electrochemical sensing applications. The composite's low energy band gap of 1.89 eV and distinct morphological features of nano sphere-shaped AgInS2 nanoparticles and rod-like structure of biochar provide a highly exposed surface area conducive to dual performance. The nanocomposite exhibited remarkable efficiency in degrading malachite green (MG), achieving 97.4% degradation in just 70 minutes, and showed excellent electrocatalytic activity in the oxidation of glufosinate herbicide, with a detection limit of 0.026 μM. Moreover, DFT studies were carried out on a simulated model material for the calculation of band gap energy values, and the consequences were in full accord with the experimental data. These results highlight the stability of the composite and efficient electroactive properties driven by the synergy between biochar and transition metals. This study opens new avenues for the development of advanced materials for environmental remediation and pollutant detection, showcasing the AgInS2@biochar nanocomposite as a versatile and effective solution for addressing various environmental challenges.

Author contributions

Firdous Ahmad Ganaie: data curation, experimentation, investigation, methodology, and original draft. Irfan Nazir: review, editing, and writing. Aaliya Qureashi: experimentation, writing, and review. Zia-ul-Haq: investigation and methodology. Arshid Bashir: review, writing, and editing. Mohsin Ahmad Bhat: supervision, review, and writing. Altaf Hussain Pandith: conceptualization, project administration, supervision, resources, review, and editing.

Conflicts of interest

Authors declare no conflicts of interest.

Data availability

The data supporting the findings of this study are available within the article (Hydrothermal synthesis of AgInS2@biochar nanocomposites for the photocatalysis and electrochemical sensing of glufosinate herbicides). Additional data are available from the corresponding author, and the data have not been submitted to any public repository.

Acknowledgements

We acknowledge the Department of Science & Technology, government of India, New Delhi, for providing facilities under DST-PURSE Programme (TPN-56945) to the Department of Chemistry, University of Kashmir, and for providing fellowship to Aaliya Qureashi under Women Scientist Scheme-A (WOS-A) [DST/WOS-A/CS-34/2021]. We are also thankful to NIT Srinagar for providing FE-SEM facility, IIT Delhi for providing TEM facility, and IISER Pune faculty for providing XPS facility for this work.

References

  1. A. M. Abdelfatah, M. Fawzy, M. E. El-Khouly and A. S. Eltaweil, Efficient Adsorptive Removal of Tetracycline from Aqueous Solution using Photosynthesized Nano-zero Valent Iron, J. Saudi Chem. Soc., 2021, 25, 101365 CrossRef CAS .
  2. A. S. Eltaweil, I. M. Mamdouh, E. M. Abd El-Monaem and G. M. El-Subruiti, Highly efficient removal for methylene blue and Cu2+ onto UiO-66 metal–organic framework/carboxylated graphene oxide incorporated sodium alginate beads, ACS Omega, 2021, 6, 23528–23541 CrossRef CAS .
  3. E. R. Sumner, A. Shanmuganathan, T. C. Sideri, S. A. Willetts, J. E. Houghton and S. V. Avery, Oxidative protein damage causes chromium toxicity in yeast, Microbiology, 2005, 151, 1939–1948 CrossRef CAS PubMed .
  4. F. A. Ganaie, Z. haq, B. Arshid, Q. Aaliya, N. Irfan, F. Kaniz, A. H. Pandith and M. A. Bhat, SnS2 decorated biochar: a robust platform for the photocatalytic degradation and electrochemical sensing of pollutants, New J. Chem., 2024, 48, 7111 RSC .
  5. S. Zinatloo-Ajabshir, M. S. Morrissey and M. Salavati-Niasari, Eco-friendly synthesis of Nd2Sn2O7-based nanostructure materials using grape juice as green fuel as photocatalyst for the degradation of erythrosine, Composites, Part B, 2019, 167, 643–653 CrossRef CAS .
  6. WHO, International Code of Conduct on the Distribution and Use of Pesticides Guidelines on Pesticide Advertising. 2010, Who/Htm/Ntd/Whopes/2010.1.
  7. L. London and J. Myers, General patterns of agrichemical usage in the southern region of South Africa, S. Afr. J. Sci., 1995, 91, 509–514 CAS .
  8. S. K. Golfinopoulos, A. D. Nikolaou, M. N. Kostopoulou, N. K. Xilourgidis, M. C. Vagi and D. T. Lekkas, Organochlorine pesticides in the surface waters of Northern Greece, Chemosphere, 2003, 50, 507–516 CrossRef CAS .
  9. Q. Aaliya, F. A. Ganaie, B. Arshid, N. Irfan, H. Zia ul, L. A. Malik, F. Kaniz, A. A. Abdullah Yahya and A. H. Pandith, Biochar waste-based ZnO materials as highly efficient photocatalysts for water treatment, J. Environ. Chem. Eng., 2022, 107256 Search PubMed .
  10. Z. ul Haq, A. H. Pandith, A. Bashir, I. Nazir and R. Tomar, Enhanced Charge Transfer and Photocatalytic Performance of Cube-Shaped Ag3PO4@Zeolite-A Nanocomposite, Mater. Chem. Phys., 2023, 302, 127701 CrossRef .
  11. S. K. Jain, M. Fazil, F. Naaz, N. A. Pandit, J. Ahmed, S. M. Alshehri, Y. Mao and T. Ahmad, Silver-doped SnO2 nanostructures for photocatalytic water splitting and catalytic nitrophenol reduction, New J. Chem., 2022, 46, 2846–2857 RSC .
  12. S. Irfan, Z. ul Haq, A. Tahir, I. Nazir, T. Iqbal, T. U. Rehman, N. Yaqoob, S. Showket, M. H. Flaifel, M. A. Shah and G. N. Dar, Multifunctional CdS/MoS2-GO Nanocomposite for Enhanced Electrochemical Dopamine Sensing and Photocatalytic Remediation, Inorg. Chem. Commun., 2025, 115003 CrossRef CAS .
  13. G. Shweta, A. Rinki, F. Mohd, M. Laxmikanta, H. Sayan, S. Sucheta, C. Biswarup, A. Tokeer, C. Chanchal and R. Manoj, Effect of Ni Metals on the [PTiW11O40]5− POM-Stabilized Self-Doped TiO2 NPs toward Visible Light-Induced Hydrogen Evolution Reactions, ACS Appl. Energy Mater., 2025, 10, 6320–6329 Search PubMed .
  14. N. A. Pandit, J. A. Ahmed and T. Ahmad, TiO2@ZrO2 p-n heterostructured nanocomposites for enhanced NO2 gas sensing, Ceram. Int., 2024, 50, 2455055–2455064 CrossRef .
  15. I. H. Lone, T. Islam, J. Ahmed, S. Ahmed, K. V. Ramanujachary and T. Ahmad, Highly Sensitive TbCrO3 Capacitive Humidity Nanosensor Prepared by Sol–Gel Combustion, ChemistrySelect, 2025, 10, e202403566 CrossRef CAS .
  16. M. Farooq, I. Nazir, H. Rafiq and M. H. Rasool, Review-Exploring Technological Innovations of Doped Rare Earth Materials, ECS J. Solid State Sci. Technol., 2023, 12(4), 47006 CrossRef .
  17. Y. F. Liu, T. Zhu, M. Deng, X. S. Tang, S. Han, A. P. Liu, Y. Z. Bai, D. R. Qu, X. B. Huang and F. Qiu, Selective and sensitive detection of copper(II) based on fluorescent zinc-doped AgInS2 quantum dots, J. Lumin., 2018, 201, 182–188 CrossRef .
  18. B. M. May, S. Parani, J. V. Rajendran and O. S. Oluwafemi, Selective detection of folic acid in the midst of other biomolecules using water-soluble AgInS2 quantum dots, MRS Commun., 2019, 9, 1306–1310 CrossRef .
  19. A. Delices, D. Moodelly, C. Hurot, Y. X. Hou, W. L. Ling, C. Saint-Pierre, D. Gasparutto, G. Nogues, P. Reiss and K. Kheng, Aqueous synthesis of DNA functionalized near-infrared AgInS2/ZnS core/shell quantum dots, ACS Appl. Mater., 2020, 12, 44026–44038 CrossRef PubMed .
  20. S. Parani and O. S. Oluwafemi, Selective and sensitive fluorescent Nano probe based on AgInS2-ZnS quantum dots for the rapid detection of Cr (III) ions in the midst of interfering ions., Nanotechnology, 2020, 31, 395501 CrossRef PubMed .
  21. A. Bahadoran, N. B. Baghbadorani, J. R. De Lile, S. Masudy-Panah, B. Sadeghi, J. Li, S. Ramakrishna, Q. Liu, B. J. Janani and A. Fakhri, Ag doped Sn3O4 nanostructure and immobilized on hyperbranched polypyrrole for visible light sensitized photocatalytic, antibacterial agent and microbial detection process, J. Photochem. Photobiol., B, 2022, 228, 112393 CrossRef PubMed .
  22. P. Ganguly, S. Mathew, L. Clarizia, S. Kumar R, A. Akande, S. J. Hinder, A. Breen and S. C. Pillai, Ternary Metal Chalcogenide Heterostructure (AgInS2–TiO2) Nanocomposites for Visible Light Photocatalytic Applications, ACS Omega, 2020, 5(1), 406–421 CrossRef .
  23. H. Tsai, J. Shaya, S. Tesana, V. B. Golovko, S. Wang, Y. Liao, C. Lu and C. Chen, Visible-Light Driven Photocatalytic Degradation of Pirimicarb by Pt-Doped AgInS2 Nanoparticles, Catalysts, 2020, 10(8), 857 CrossRef .
  24. A. Malankowska, D. Kulesza, J. Sowik, O. Cavdar, T. Klimczuk and G. T. Medynska, The Effect of AgInS2, SnS, CuS2, Bi2S3 Quantum Dots on the Surface Properties and Photocatalytic Activity of QDs-Sensitized TiO2 Composite, Catalysts, 2020, 10(4), 403 CrossRef CAS .
  25. T. Li, H. He, P. Zhang, X. Zhao, W. Yin and X. Tu, The synergy of step-scheme heterojunction and sulfur vacancies in AgInS2/AgIn5S8 for highly efficient photocatalytic degradation of oxy tetracycline, Colloids Surf., A, 2022, 128946 CrossRef CAS .
  26. Y. Li, Y. Liu, G. Gao, Y. Zhu, D. Wang, M. Ding, T. Yao, M. Liu and W. Sheng, L-cysteine and urea synergistically-mediated one-pot one-step self-transformed hydrothermal synthesis of p-Ag2S/n-AgInS2 core-shell heteronano flowers for photocatalytic MO degradation, Appl. Surf. Sci., 2021, 149279 CrossRef CAS .
  27. M. Mousavi-Kamazani and M. Salavati-Niasari, A simple microwave approach for synthesis and characterization of Ag2S-AgInS2 nanocomposites, Composites, Part B, 2014, 56, 490–496 CrossRef CAS .
  28. M. A. Franzman and R. L. Brutchey, Solution-phase synthesis of well-defined indium sulfide nanorods, Chem. Mater., 2009, 21, 1790–1792 CrossRef CAS .
  29. M. Rahman, S. Shaheen and T. Ahmad, Photocatalytic transformation of organic pollutants and remediation strategies of carbon emissions and nitrogen fixation in inland water, Mater. Today Catal., 2025, 9, 100103 Search PubMed .
  30. J. Fito and T. T. I. Nkambule, Synthesis of biochar-CoFe2O4 nanocomposite for adsorption of methyl paraben from wastewater under full factorial experimental design, Environ. Monit. Assess., 2023, 241 CrossRef CAS .
  31. R. Gayathri, K. P. Gopinath and P. S. Kumar, Adsorptive separation of toxic metals from aquatic environment using agro waste biochar: application in electroplating industrial wastewater, Chemosphere, 2021, 262, 128031 CrossRef CAS .
  32. R. V. Hemavathy, P. S. Kumar, K. Kanmani and N. Jahnavi, Adsorptive separation of Cu(II) ions from aqueous medium using thermally/chemically treated Cassia fistula based biochar, J. Cleaner Prod., 2020, 249, 119390 CrossRef CAS .
  33. L. A. Malik, A. Bashir, T. Manzoor and A. H. Pandith, Microwave-assisted synthesis of glutathione-coated hollow zinc oxide for the removal of heavy metal ions from aqueous systems, RSC Adv., 2019, 9, 15976–15985 RSC .
  34. A. Bashir, L. A. Malik, S. Ahad, T. Manzoor, M. A. Bhat, G. N. Dar and A. H. Pandith, Removal of heavy metal ions from aqueous system by ion-exchange and bio sorption methods, Environ. Chem. Lett., 2019, 729–754 CrossRef CAS .
  35. A. Bashir, A. H. Pandith, L. A. Malik, A. Qureashi, F. A. Ganaie and G. N. Dar, Magnetically recyclable L-cysteine capped Fe3O4 Nano adsorbent: A promising pH guided removal of Pb(II), Zn(II) and HCrO4-contaminants, J. Environ. Chem. Eng., 2021, 9, 105880 CrossRef CAS .
  36. L. Zhong, R. Ni, L. Zhang, Z. He, H. Zhou and L. Li, Determination of total arsenic in soil by gas chromatography after pyrolysis, Microchem. J., 2019, 146, 568–574 CrossRef CAS .
  37. L. C. Soliman and K. K. Donkor, Micellar electrokinetic chromatography method development for simultaneous determination of thiabendazole, carbendazim, and flubendazole, J. Environ. Sci. Health, Part B, 2014, 49(3), 153–158 CrossRef CAS .
  38. S. Grujic, M. Radisic, T. Vasiljevic and M. Lausevic, Determination of carbendazim residues in fruit juices by liquid chromatography-tandem mass spectrometry, Food Addit. Contam., 2005, 22(11), 1132–1137 CrossRef CAS .
  39. N. Pourreza, S. Rastegarzadeh and A. Larki, Determination of fungicide carbendazim in water and soil samples using dispersive liquid-liquid microextraction and microvolume UV-vis spectrophotometry, Talanta, 2015, 134, 24–29 CrossRef CAS PubMed .
  40. L. F. Lopez, A. G. Lopez and M. V. Riba, HPLC Method for Simultaneous Determination of Fungicides: Carbendazim, Metalaxyl, Folpet, and Propiconazole in Must and Wine, J. Agric. Food Chem., 1989, 37(3), 684–687 CrossRef CAS .
  41. I. Nazir, A. Qureashi, A. Bashir, Z. U. Haq, F. A. Ganaie, G. N. Dar and A. H. Pandith, Synergistic Sb2S3-NiS2 Heterostructure: A Robust Electrocatalyst for Electrochemical Sensing of Hg (II), As (III) Ions and Carbendazim Fungicide, J. Environ. Chem. Eng., 2024, 12(3), 112793 CrossRef CAS .
  42. I. Nazir, Z. U. Haq, A. Bashir, A. Qureashi, F. A. Ganaie, K. Fatima, S. Irfan, G. N. Dar and A. H. Pandith, Electrochemical Detection and Catalytic Reduction of Nitrobenzene Using a Bimetallic NiS2/Fe3S4 Magnetic Heterostructure: An Innovative Approach for Environmental Remediation, New J. Chem., 2024, 48(11), 4909–4921 RSC .
  43. Z. U. Haq, A. Qureashi, I. Nazir, F. A. Ganaie, A. Bashir, L. A. Malik and A. H. Pandith, Fabrication of 2D Graphene Oxide Incorporating S-Scheme Sn2S3-In2S3 Heterojunctions for Enhanced Photocatalytic Mineralization of Organic Pollutants, New J. Chem., 2024, 48, 5988–5999 RSC .
  44. H. Li, S. Weizhe, C. Xingqiang, L. Yanhui, H. Baorong, Z. Xiaoping, W. Yuqi, C. Lianjun, Z. Pengfei and L. Junru, AgInS2 and graphene co-sensitized TiO2 photoanodes for photocathodic protection of Q235 carbon steel under visible light, Nanotechnology, 2020, 31, 305704 CrossRef CAS PubMed .
  45. M. Hashemkhani, M. Loizidou, A. J. MacRobert and H. Y. Acar, One-Step Aqueous Synthesis of Anionic and Cationic AgInS2 Quantum Dots and Their Utility in Improving the Efficacy of ALA-Based Photodynamic Therapy, Inorg. Chem., 2022, 61(6), 2846–2863 CrossRef CAS .
  46. K. Mansaray and A. Ghaly, Thermal degradation of rice husks in nitrogen atmosphere, Bioresour. Technol., 1998, 65, 13–20 CrossRef CAS .
  47. L. Zhao, X. Cao, O. Mašek and A. Zimmerman, Heterogeneity of biochar properties as a function of feedstock sources and production temperatures, J. Hazard. Mater., 2013, 256–257, 1–9 CAS .
  48. A. Karthika, V. Ramasamy Raja, P. Karuppasamy, A. Suganthi and M. Rajarajan, A novel electrochemical sensor for the determination of hydroquinone in water using FeWO4/SnO2 nanocomposite immobilised modified glassy carbon electrode, Arabian J. Chem., 2020, 13(2), 4065–4081 CrossRef CAS .
  49. A. Karthika, S. Selvarajan, P. Karuppasamy, A. Suganthi and M. Rajarajan, A novel highly efficient and accurate electrochemical detection of poisonous inorganic Arsenic (III) ions in water and human blood serum samples based on SrTiO3/β-cyclodextrin composite, J. Phys. Chem. Solids, 2019, 127, 11–18 CrossRef CAS .
  50. L. Mohanty, D. S. Pattanayak and S. K. Dash, An efficient ternary photocatalyst Ag/ZnO/g-C3N4 for degradation of RhB and MG under solar radiation, J. Indian Chem. Soc., 2021, 98, 100180 CrossRef CAS .
  51. N. Arsalani, S. Bazazi, M. Abuali and S. Jodeyri, A new method for preparing ZnO/CNT nanocomposites with enhanced photocatalytic degradation of malachite green under visible light, J. Photochem. Photobiol., A, 2020, 389, 112207 CrossRef CAS .
  52. Y. Yulizar, D. O. B. Apriandanu and R. I. Ashna, La2CuO4-decorated ZnO nanoparticles with improved photocatalytic activity for malachite green degradation, Chem. Phys. Lett., 2020, 755, 137749 CrossRef CAS .
  53. T. T. Pham, L. Lieu, A. T. Nguyen and T. T. Thuy Le, Chitosan films integrated with zinc oxide/reduced graphene oxide nanocomposite for malachite green degradation and antibacterial activity in sustainable packaging, Int. J. Biol. Macromol., 2025, 318, 145009 CrossRef CAS .
  54. P. B. Sreelekshmi, R. R. Pillai, B. Binish and A. P. Meera, Enhanced Photocatalytic Degradation of Malachite Green Using Highly Efficient Copper Oxide/Graphene Oxide Nanocomposites, Top Catal., 2022, 65, 1885–1898 CrossRef CAS .
  55. S. S. Hegde, R. S. C. Bose, B. S. Surendra, S. Vinoth, P. Murahari and K. Ramesh, SnS-Nanocatalyst: Malachite green degradation and electrochemical sensor studies, Mater. Sci. Eng., B, 2022, 283, 115818 CrossRef CAS .
  56. M. F. Elkady and H. S. Hassan, Photocatalytic Degradation of Malachite Green Dye from Aqueous Solution Using Environmentally Compatible Ag/ZnO Polymeric Nanofibers, Polymers, 2021, 13, 2033 CrossRef CAS PubMed .
  57. F. F. Alharbi, S. Gouadria, M. Abdullah, A. G. Abid, M. Nisa, M. T. Farid and S. Aman, CuO/ZnTe nanocomposite for photodegradation of malachite green from industrial effluents to clean environment, J. Mater. Sci.: Mater. Electron., 2023, 34, 2150 CrossRef CAS .
  58. R. B. Mampilly, A. Pathan and C. P. Bhasin, Visible light-assisted degradation of malachite green Dye Using waste tea-mediated zinc nanoparticles, Int. J. Thin Film Sci. Technol., 2023, 12, 39–51 CrossRef .
  59. L. Shi, W. Si, F. Wang and W. Qi, Construction of 2D/2D layered g-C3N4/Bi12O17Cl2 hybrid material with matched energy band structure and its improved photocatalytic performance, RSC Adv., 2018, 8, 24500–24508 RSC .
  60. X. Fang, S. Yang, K. Chingin, L. Zhu, X. Zhang, Z. Zhou and Z. Zhao, Quantitative Detection of Trace Malachite Green in Aquiculture Water Samples by Extractive Electrospray Ionization Mass Spectrometry, Int. J. Environ. Res. Public Health, 2016, 13, 814 CrossRef PubMed .
  61. X. Wu, G. Zhang, Y. Wu, X. Hou and Z. Yuan, Simultaneous determination of malachite green, gentian violet and their leuco-metabolites in aquatic products by high-performance liquid chromatography-linear ion trap mass spectrometry, J. Chromatogr. A, 2007, 1172, 121–126 CrossRef CAS .
  62. Z. Fengyuan, Z. Jing and Z. Zhengzhu, Selective detection of glufosinate using CuInS2 quantum dots as a fluorescence probe, RSC Adv., 2017, 7, 48077 RSC .
  63. S. Mengmeng, Z. Lying, L. Tau, L. Zhiwei, S. Gehong, W. Chun, S. Chang, D. Rui, H. Mingxia, R. Hanbing and W. Yanying, Construction of CuO/Fe2O3 Nanozymes for Intelligent Detection of Glufosinate and Chlortetracycline Hydrochloride, ACS Appl. Mater. Interfaces, 2023, 15(47), 54466–54477 CrossRef .
  64. S. Nafisah, M. Morsin, N. A. Jumadi, N. Nayan, N. S. M. Shah, N. L. Razali and N. Z. An’nisa, Improved Sensitivity and Selectivity of Direct Localized Surface Plasmon Resonance Sensor Using Gold Nano bipyramids for Glyphosate Detection, IEEE Sens. J., 2020, 20, 2378–2389 CAS .
  65. B. Bucur, F. D. Munteanu, J. L. Marty and A. Vasilescu, Advances in enzyme-based biosensors for pesticide detection, Biosensors, 2018, 8, 27 CrossRef PubMed .
  66. A. Khenifi, Z. Derriche, C. Forano, V. Prevot, C. Mousty, E. Scavetta, B. Ballarin, L. Guadagnini and D. Tonelli, Glyphosate and glufosinate detection at electrogenerated NiAl-LDH thin films, Anal. Chim. Acta, 2009, 654, 97–102 CrossRef CAS PubMed .

This journal is © The Royal Society of Chemistry and the Centre National de la Recherche Scientifique 2025
Click here to see how this site uses Cookies. View our privacy policy here.