Aaliya
Qureashi
a,
Altaf Hussain
Pandith
*a,
Arshid
Bashir
a,
Lateef Ahmad
Malik
a,
Taniya
Manzoor
a,
Faheem A.
Sheikh
b,
Kaniz
Fatima
a and
Zia-ul
Haq
a
aLaboratory of Nanoscience and Quantum Computations, Department of Chemistry, University of Kashmir, Hazratbal, Srinagar, J&K, India. E-mail: altafpandit23@gmail.com; Fax: +91-194-2414049; Tel: +91-194-2424900, +91-7006429021
bDepartment of Nanotechnology, University of Kashmir, Srinagar-190006, Kashmir, India
First published on 5th January 2023
Glyphosate [N-(phosphonomethyl)glycine] is a widely used phosphonate herbicide for different agricultural purposes. Due to its widespread use, suspected toxicity, and ubiquitous bioaccumulation, it is one of the most harmful contaminants found in drinking water. This demands efficient sensing and removal of glyphosate from contaminated water. Here, we report the decoration of novel and highly porous biochar with nanozero-valent iron (nZVI) nanoparticles to develop an efficient electrochemical sensor for the trace detection of glyphosate. The as-synthesized composite was thoroughly characterized by various state-of-the-art instrumental techniques. The electron micrographs of the composite materials revealed the cavity-like structure and the abundant loading of nZVI nanoparticles. FTIR and XPS analyses confirmed the presence of oxygen-rich functionalities and Fe(0) in the composite nanostructure. Electrochemical analysis through CV, LSV, and DPV techniques suggested efficient sensing activity with a limit of detection as low as 0.13 ppm. Furthermore, the chronopotentiometric response suggested excellent and superior stability for long-term applications. To gain more insight into the interaction between glyphosate and the composite material, DFT calculations were carried out. The Frontier Molecular Orbital study (FMO), Molecular Electrostatic Potentials (MEPs), and Density of States (DOS) suggest an increase in the electron density, an increase in the DOS, and a decrease in the HOMO–LUMO band gap by combining nZVI nanoparticles and biochar. The results suggest more facile electron transfer from the composite for trace detection of glyphosate. As a proof of concept, we have demonstrated that real-time analysis of milk, apple juice, and the as-synthesized composite shows promising results for glyphosate detection with an excellent recovery rate.
Several methods such as high-performance liquid chromatography,5 capillary electrophoreses,6 and gas chromatography7,8 have been used to detect glyphosate. Despite their selectivity and high sensitivity, some shortcomings such as high cost and sample pre-treatment limit their use on a large scale.9 Keeping this in mind, the development of electrochemical sensors has grabbed attention as they offer advantages in terms of a low cost, simple mode of operation, great sensitivity, specificity, and selectivity.10 It is noteworthy to mention here that glyphosate is not electrochemically active; it becomes difficult to monitor glyphosate without modification of electrodes. In this regard, various materials such as MOFs,11 chitosan,12 layered double hydroxides,13 gold nanoparticles,14 zeolites,15 CNTs,16 and carbon-based electrodes17 were reported in the literature for the electrochemical monitoring of glyphosate.
Biochar is a porous carbonaceous material obtained from the pyrolysis of different types of biomass. It has a strong prospect for its use in the abatement of environmental toxicants in wastewater management. The use of biochar offers several advantages such as a high surface area with a porous structure, less toxicity, abundant functional groups, and redox activity.18 However, biochar-based materials have the limitations of less recovery and generation of secondary sludge. Therefore, there is emerging interest in the use of easily recoverable magnetically active materials owing to their superior properties such as magnetization, a large surface area with a strong affinity for contaminants, and easy recovery. So far, several magnetic materials have been continuously used for tuning the surface of biochar to enhance its surface charge capacity for better interaction and adsorption of organophosphorus pesticides.19–21 In this arena of research, nano-zerovalent iron (nZVI) nanoparticles are the emerging magnetic material of choice due to their cost-effective nature, abundance, and active surface area with high reactivity. The nanoscale zero-valent iron (nZVI) is highly reactive toward the majority of pollutants such as dyes, antibiotics, heavy metal ions, and halogenated compounds.22–27
Under batch experimental conditions, nanoparticles are susceptible to aerial oxidation and agglomeration, which decreases the surface activity of the nanoparticles.28 Among others, the impregnation of nZVI nanoparticles into porous material is ideal for the uptake of pollutants.
Herein, we report the synthesis of a nZVI@Biochar nanocomposite by adopting a simple ultra-sonication method. Various characterization techniques were employed to analyze and characterize the prepared magnetic biochar. The capacitive properties of nZVI and the prepared nanocomposites were inferred from magnetic studies. The as-synthesized magnetic biochar was found to be an efficient electrochemical sensor for glyphosate. The electrochemical response of the prepared sensor (nZVI@Biochar) for glyphosate was analyzed by utilizing cyclic voltammetry (CV), differential pulse voltammetry (DPV), and linear sweep voltammetry (LSV). The electrochemical sensing of glyphosate was further supported by DFT studies and the results were in complete harmony with each other.
For the synthesis of the nZVI@Biochar nanocomposite, 1.6 g of biochar was dispersed in a three-necked flask containing 100 mL of 0.8 g L−1 FeSO4·7H2O. For efficient adsorption of Fe2+ ions on biochar, the solution was continuously ultra-sonicated for eight hours. Thereafter, 50 mL of 0.2 M aqueous NaBH4 was added dropwise into the flask to ensure the complete reduction of Fe2+ to Fe0 under continuous N2 gas purging. The reduction process involved in this step is according to Eqn (1). Moreover, the metal ions naturally present in the biochar such as Mg, Ca, Si, and P also undergo a reduction in the presence of sodium borohydride apart from zerovalent iron (nZVI). The as-synthesized magnetic biochar (nZVI@Biochar) nanocomposite was removed by applying an external magnetic retriever and was washed several times with ethanol and deionized water and finally was oven-dried at 30 °C (Scheme 1).
2Fe3+ (aq) + 6BH4− (aq) + 18H20 → 2Fe0(s) + 6B(OH)3 (aq) + 21H2 (aq) | (1) |
The XRD spectra of biochar show a diffraction peak at 2θ = 27° which is indicative of the amorphous nature of biochar. The sharp diffraction peak at 2θ = 45° assigned to the 110 plane in the nZVI@Biochar nanocomposite is due to the presence of face-centered α-Fe0 particles, which indicates successful coating of nZVI on biochar (see Fig. 1b). Moreover, the other less intense peaks present in the XRD spectrum are probably due to iron oxide impurities on the surface of biochar, as reported in the literature.30 The thermograms of the as-synthesized materials nZVI and nZVI@Biochar are shown in Fig. 1c, where it is evident that in the case of nZVI there is a small weight loss which is due to the adsorbed moisture over the surface of nZVI. Compared with the nZVI@Biochar nanocomposite thermogram, it exhibits a two-stage weight-loss behavior. At first, there was not any significant weight loss up to 240 °C. After 240 °C and up to 300 °C, there was a slight drop in the weight of around 7%, which is attributed to the decomposition of volatile matter. Subsequently, a second weight loss of about 50% was observed from 390 °C to 560 °C due to the thermal degradation of hemicellulosic and cellulosic content present in the biochar. Above 560 °C, no further weight loss was observed which indicates the complete carbonization and decomposition of the different units of biochar present in the nZVI@Biochar nanocomposite. Therefore, it can be concluded that the composite (nZVI@Biochar) contains up to 30% nZVI, which proves the successful modification of biochar.
The magnetic properties of the as-prepared materials were determined by Vibrating Sample Magnetometry at room temperature (see Fig. 1d). As can be seen from the figure, the saturation magnetization of nZVI was around 27.93 emu g−1 and for the prepared nanocomposite (nZVI@Biochar) it was 17 emu g−1. The rationale for the decrease in the saturation magnetization value of nZVI@Biochar was due to the presence of a non-magnetic biochar component. Other such similar reports are available in the literature.30
The structural integrity of the nanocomposite material was investigated using XPS analysis (Fig. 2a). The main composition of the prepared nanocomposite is C, O, and Fe with a small amount of sulfur. The C 1s spectra of nZVI@Biochar can be deconvoluted into three peaks at 284.7, 286.90, and 287.2 eV, which can be attributed to the presence of C–C and C–O bonds, respectively (Fig. 2b). Likewise, the O 1s spectra show two peaks around 530.48 and 528.26 eV, which indicate the presence of various oxygen-containing functional groups (C–O, C–OOH, and CO) respectively, and the 527.99 eV peak is due to the presence of an Fe–O bond (Fig. 2c). For iron Fe 2p spectra, the specific binding energy peaks at 712, 709, and 707 eV were ascribed to the presence of Fe3+, Fe2+ and Fe0, respectively (Fig. 2d). The data are in agreement with those reported in the literature.31
Fig. 2 (a) Compete range XPS, (b) C 1s XPS, (c) O 1s XPS and (d) Fe 2p spectra of the nZVI@Biochar composite. |
Lotus, which is botanically called Nelumbonucifera, possesses an underwater stem that maintains air pockets throughout the length. The lateral cross-section, as shown in Fig. 3a and b, shows a wheel-like structure with pores in a symmetrical fashion, and the horizontal cross-section (see Fig. 3c and d) shows that the pores extend throughout the length of the stem, which is justified by FE-SEM micrographs. This porous nature of the lotus stem can be exploited as a raw material for designing high surface area and porous materials for a wide range of applications. In our case, we used it as a raw material for synthesizing a biochar-zerovalent composite, keeping in view the porous nature that could incorporate the metal as well as providing a high surface area for analyte interaction. However, it is possible only if the porous nature and cavity-type structure of the lotus stem are maintained even after the event of initial treatment through annealing. To get an insight into that, we carried out the FE-SEM analysis of the material after annealing and incorporation of metal.
Fig. 3 Lateral and horizontal cross-sections of the lotus stem describing the porous cavity throughout the length of the stem. |
The surface morphology of nZVI and the nZVI@Biochar nanocomposite was examined by scanning electron microscopy. From the FE-SEM micrographs, we can see the monodisperse spherical particles of nZVI with particle sizes in the 100 nm range (see Fig. 4a and b). The FE-SEM micrographs of porous bare biochar show cavity-like structures within the biochar matrix (see Fig. 4c and d). We observed that the biochar retained the cavity and tubular-like structures of its parent lotus stem biomass material, during pyrolysis.
Fig. 4 FE-SEM micrograph analysis of (a and b) the as-synthesized nZVI particles and (c and d) porous surface of biochar. |
Furthermore, the tubular cavities in the biochar are expected to offer an ideal environment for decorating nZVI nanoparticles and the adsorption and interaction of contaminants such as glyphosate (see Fig. 5a and b). The impregnation of nZVI nanoparticles within the biochar cavity prevents the agglomeration of these nZVI particles (Fig. S1, ESI†). Moreover, TEM micrographs manifested the successful loading of nZVI onto the biochar surface with less agglomeration of nZVI particles (see Fig. 5c and d). To determine the composition of the as-synthesized material, we have carried out elemental analysis as shown in Fig. S2, ESI.† As can be seen from the elemental mapping, the final synthesized material possesses 30.3 wt% iron with nearly uniform distribution over the composite.
Fig. 5 FE-SEM micrograph analysis of (a and b) the nZVI@Biochar composite confirming loading of nZVI on the biochar surface, and (c and d) TEM micrographs of the nZVI@Biochar composite. |
Besides, the effect of pH ranging from 2–9 on the behavior of the nZVI@Biochar electrode was examined (see Fig. 6b). There was no significant shift in the reduction peak of glyphosate with an increase in the pH. However, the maximum current intensity was observed at pH = 7. This may be due to the deprotonation of glyphosate functional groups that resulted in its increase in complexation potential with the iron metal cluster which consequently enhances the sensing response of the nZVI@Biochar electrode.
After the initial analysis with cyclic voltammetry (see Fig. S3, ESI†), we carried out linear sweep voltammetry (LSV) with varying concentrations of glyphosate in the electrolyte on the nZVI@Biochar modified electrode. As can be seen from Fig. 7, the peak current increases with an increase in the concentration of glyphosate in the electrolyte, confirming that the reduction peak is due to glyphosate. The peak current shows a linear relationship with a concentration range from 2–8 ppm. This linear response prompted us to analyze this material further for glyphosate sensing. To carry out the sensing efficiency of nZVI@Biochar for glyphosate, we carried out differential pulse voltammetry (DPV) at different concentrations of glyphosate in the electrolyte. As expected, the peak current increases with increasing concentration of glyphosate and an increase in current demonstrated a linear response, which is a prerequisite for any sensor. The critical parameters for any sensing materials are sensing efficiency denoted by the limit of detection and its stability. We have calculated the LOD from the current vs. concentration response in DPV analysis which was determined to be 0.13 ppm. Both LSV and DPV analyses demonstrate the efficient sensing capability of nZVI@Biochar towards glyphosate.
In addition to this, the limit of detection of the nZVI@Biochar sensor towards glyphosate was compared with other sensors already reported in the literature as shown in Table 1.
S. no. | Sensors | Method | Limit of detection (LOD) | References |
---|---|---|---|---|
1 | CuAl-LDH/Gr nanocomposite | Stripping voltammetry | 1 × 10−9 M | 13 |
2 | Graphite powder/mineral oil paste electrode | Square wave voltammetry | 2 × 10−9 M | 32 |
3 | Cu-BTC MOF | Differential pulse voltammetry | 1.4 × 10−13 M | 11 |
4 | Carbon paste electrode-modified biochar copper(II) hexadecafluoro-29H,31 phthalocyanine complex | Square wave voltammetry | 2 × 10−8 M | 33 |
5 | Organophilic silane in the interlayer space of acid-treated smectite | Square wave voltammetry | 9.8 × 10−7 M | 34 |
6 | Horseradish peroxidase polymeric matrix | Cyclic voltammetry | 1.7 × 10−6 M | 35 |
7 | Copper nanoparticles and reduced graphene oxide | Differential pulse voltammetry | 1.9 × 10−7 M | 36 |
8 | Cu-BTC MOF/g-C3N4 | Photocurrent | 1.3 × 10−13M | 37 |
9 | Graphite oxide sensor | Square wave voltammetry | 1.7 × 10−8M | 38 |
10 | nZVI@Biochar | Differential pulse voltammetry | 7 × 10−7 M (0.13 ppm) | This work |
After optimizing the geometries of each of the materials, we calculated the interaction energies of glyphosate with biochar and the composite. As for the electrochemical transformations, favorable interaction between glyphosate and the material is a prerequisite. The interaction energies of glyphosate with both biochar and composite materials were calculated based upon optimized geometries. The interaction energy of the energetically favorable configuration of glyphosate-biochar (Egly-BC), nZVI@Biochar (EBC-nZVI), and nZVI@Biochar-glyphosate (Egly-BC-nZVI) was calculated by using the following equation shown:
ΔE = Ecomplex − [Egly-BC + Egly-BC-nZVI] | (2) |
Moreover, to validate our results and for higher accuracy, we have calculated basis set superposition error (BSSE)41 by employing eqn (3)
(3) |
The calculated interaction energy value of glyphosate biochar was found to be −598.614 kJ mol−1 and its counterpoise interaction energy values were found to be −631.9965 kJ mol−1. Similarly, the interaction energy value for nZVI@Biochar-glyphosate was found to be −1417.77 kJ mol−1 and its corrected counterpoise interaction energy was determined to be −1575.3 kJ mol−1. Therefore, the basis set superposition error was found to be −33.3825 kJ mol−1 and −157.53 kJ mol−1 for glyphosate-biochar and nZVI@Biochar-glyphosate, respectively. This error is found to be very small, which is indicative of the accuracy of our calculated results.
In addition to this, the interaction energy of glyphosate-biochar got enhanced with the incorporation of nZVI in the complex, predicting better chemical interaction between the biochar-glyphosate functional moieties with Fe(0). To confirm these interactions, we have carried out theoretical IR (Fig. 9d), which is in line with IR carried out experimentally. In addition to this, the preferential site of interactions in the composite material was found to be the feat (positon-16) of nZVI and oxygen at position 24 (O-24) of glyphosate. Single point energy (SPE) scans were carried out to find the bond distance between Fe-16 nZVI and O-24 at which the interaction is maximum (Fig. 9e). All of these suggest a favorable interaction between Fe and glyphosate, which is not the case with the biochar.
Apart from interaction, the critical factor that decides electron transfer between the electrode and analyte is the HOMO–LUMO energy gap of biochar-glyphosate and nZVI@Biochar-glyphosate systems. The lower the HOMO–LUMO gap the smaller the energy barrier for electron transfer (the activation overpotential) between the electrode and analyte in the electrolyte. To get an insight into this, we explored the frontier molecular orbitals (FMOs) and performed molecular electrostatic potential (MEP) calculations upon the interaction of biochar and nZVI@Biochar with the glyphosate system (see Fig. 10a). The calculated band gap energy values and other possible HUMO–LUMO orientations shown in (T2, and Fig. S4, ESI†) demonstrate that the band gap decreases noticeably by the incorporation of zero-valent Fe into the biochar-glyphosate system. Furthermore, we also calculated the band gap energy values of Fe(0), Fe2O3, and Fe3O4 with glyphosate (Fig. S5, ESI†). The band gap energy value was found to be more in the case of the Fe2O3-glyphosate and Fe3O4-glyphosate systems in comparison to Fe(0)-glyphosate, which is not suitable for feasible electron transfer. From these observations, we can also conclude that Fe(0) plays a major role in the electrochemical catalysis of glyphosate.
To verify the results, we calculated the density of states (DOS) in the biochar-glyphosate composite with and without nZVI. In the case of biochar-glyphosate, the density of states is low, but the incorporation of nZVI in the composite results in a sudden jump in the DOS (see spectra in Fig. 10b). It is known that an increase in the DOS results in a decrease in the band gap, which is a direct representation of a reduction in the energy gap. All these results suggest facile electron transport between glyphosate and the nZVI@Biochar composite, which could be exploited for designing an efficient sensing platform for glyphosate in aqueous media. This positive interaction between the nZVI@Biochar composite and glyphosate is also supported by a low HOMO–LUMO gap and high DOS, thus strengthening the results obtained through CV and DPV (Fig. 6 and 7).
The same is further supported by MEP, which allows us to visualize the variable charge distribution within the molecule. The red regions indicate electron-rich regions, whereas the green regions indicate neutral areas. As is evident from Fig. 10c with the incorporation of nZVI in GLY-BC (Glyphosate-Biochar), the red region around the biochar increases compared to the parent biochar materials on account of the presence of iron in the biochar indicating a negative electrostatic potential. The increase in the electron density due to the incorporation of nZVI results in higher reducing power of the composite towards the glyphosate, which is in line with the electrochemical data.
Taken together, both experimental and theoretical calculations suggest better electrochemical reduction dynamics compared to the biochar alone. In the case of nZVI@Biochar, the Fe center acts as a mediator as no reduction peak was observed with biochar alone (see Fig. 6a). Since it is an inner sphere electron transfer reaction, the analyte is required to have a favorable interaction with the electrocatalyst. As mentioned above, the interaction energy of glyphosate-biochar got enhanced with the incorporation of nZVI in the complex, predicting better chemical interaction between the biochar-glyphosate functional moieties and Fe(0).
With this, we came to a conclusion why nZVI@Biochar is better suited compared to biochar alone. Secondly, the inner sphere electrochemical transformations occur via interaction between the analyte and catalytic site on the electrocatalyst. The preferential site of interactions in the composite material was found to be Fe at (positon-16) of nZVI and oxygen at position 24 (O-24) of glyphosate as discussed above. This is in line with our experimental results that suggest a facile electrochemical charge transfer incorporation of nZVI on biochar. Thus, nZVI acts as both an interaction site and an electrocatalytic reaction/sensing site.
Fig. 11 Linear sweep voltammograms of spiked samples of milk and juice with a known amount of glyphosate. |
Glyphosate (real) | Before spiking (unspiked) | After spiking (spiked) | Recovery factor | |
---|---|---|---|---|
Milk | 0.0105 mA | 0.0102 mA | 0.0204 mA | 97% |
Juice | 0.0105 mA | 0.0106 mA | 0.0212 mA | 101% |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2na00610c |
This journal is © The Royal Society of Chemistry 2023 |