Next Article in Journal
Adsorption and Recombination of H+ and H3O+ on Graphene-Supported Pt1, Pt13, and Pt14 Nanoclusters: A First Principles Study
Previous Article in Journal
Basic Properties of ZnO, Ga2O3, and MgO—Quantitative IR Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermally Synthesized MoS2 as Electrochemical Catalyst for the Fabrication of Thiabendazole Electrochemical Sensor and Dye Sensitized Solar Cells

1
Department of Chemistry, M.M.D.C, Moradabad, M.J.P. Rohilkhand University, Bareilly 244001, India
2
School of Materials Science and Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
3
Department of Materials Science and Engineering, WW4-LKO, University of Erlangen-Nuremberg, Martensstrasse 7, 91058 Erlangen, Germany
4
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
5
Faculdade de Engenharia, Universidade Lusófona—Centro Universitário de Lisboa, Campo Grande, 376, 1749-024 Lisbon, Portugal
6
Centro de Química Estrutural, Institute of Molecular Sciences, Instituto Superior Técnico, Universidade de Lisboa, Av. Rovisco Pais, 1049-001 Lisbon, Portugal
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Catalysts 2024, 14(2), 107; https://doi.org/10.3390/catal14020107
Submission received: 4 December 2023 / Revised: 18 January 2024 / Accepted: 23 January 2024 / Published: 26 January 2024
(This article belongs to the Section Catalytic Materials)

Abstract

:
In this work we reported the hydrothermal preparation of molybdenum disulfide (MoS2). The phase purity and crystalline nature of the synthesized MoS2 were examined via the powder X-ray diffraction method. The surface morphological structure of the MoS2 was examined using scanning electron microscopy and transmission electron microscopy. The specific surface area of the MoS2 was calculated using the Brunauer-Emmett-Teller method. The elemental composition and distribution of the Mo and S elements were determined using energy-dispersive X-ray spectroscopy. The oxidation states of the Mo and S elements were studied through employing X-ray photoelectron spectroscopy. In further studies, we modified the active surface area (3 mm) of the glassy carbon (GC) electrode using MoS2 as an electrocatalyst. The MoS2 modified GC electrode (MSGC) was used as an electrochemical sensor for the detection of thiabendazole (TBZ). Linear sweep voltammetry (LSV) was used as the electrochemical sensing technique. The MSGC exhibited good performance in the detection of TBZ. A limit of detection of 0.1 µM with a sensitivity of 7.47 µA/µM.cm2 was obtained for the detection of TBZ using the LSV method. The MSGC also showed good selectivity for the detection of TBZ in the presence of various interfering compounds. The obtained results showed that MoS2 has good electrocatalytic properties. This motivated us to explore the catalytic properties of MoS2 in dye sensitized solar cells (DSSCs). Thus, we have fabricated DSSCs using MoS2 as a platinum-free counter electrode material. The MoS2 counter electrode-based DSSCs showed good power conversion efficiency of more than 5%. We believe that the present work is beneficial for the scientific community, and especially for research surrounding the design and fabrication of catalysts for electrochemical sensing and DSSC applications.

Graphical Abstract

1. Introduction

At present, the energy crisis, environmental pollution, and food safety related issues are major concerns which demand extensive attention [1,2,3]. Ensuring the safety of food is a critical priority for public health, especially when it comes to agricultural products that play a vital role in providing essential nutrition to the population [4]. Although there has been a historical emphasis on the safety of veterinary drugs, recent research has highlighted a new potential hazard: the existence of various pesticides in crops cultivated in soil [5]. The application of pesticides on crops is a vital element in cultivating agricultural goods, providing advantages such as enhanced product quality [6,7,8]. However, if pesticides are applied incorrectly or excessively, it can lead to the accumulation of pesticide residues, which pose a substantial threat to global environmental safety and public health [9]. Prolonged exposure to these pesticides has been associated with various adverse health effects, encompassing neurodegenerative, respiratory, reproductive, and developmental toxicity [10]. Of particular concern are the effects observed during crucial developmental phases, which can have enduring repercussions on individuals, especially concerning brain development and the endocrine system [11]. A notable harmful pesticide is the fungicide and insecticide thiabendazole (TBZ), extensively used in agriculture to combat mildew and blight on crops [12]. TBZ residues have been identified in both crops and farm runoff, and its biological activity and persistence enhance its potential to interfere with thyroid hormone balance, induce liver damage, and potentially lead to severe cases of cancer [10]. Consequently, it is essential to establish accurate and dependable methods for identifying and measuring TBZ in food matrices [9]. Ensuring the safety of agricultural products mandates thorough pesticide residue testing before the products enter the market [13]. Several effective assessment techniques have been documented for identifying pesticide residues in fruits, encompassing chromatographic methods, spectrometric approaches, and immunoassays [13,14,15,16]. However, these methods often necessitate advanced instruments, costly reagents, and extensive sample preparation, constraining their suitability for on-the-spot detection and quick screening [17,18,19]. In contrast, electrochemical sensors offer a straightforward, cost-efficient, and rapid alternative method for the detection of pesticide residues in food samples [20,21,22]. In recent times, numerous electrochemical sensors have emerged as candidates for determining the presence of TBZ and other pesticide residues in fruits, utilizing various nanomaterials and sensing strategies [18]. These encompass sensors based on graphene, metal oxide nanocomposites, conducting polymers, and more. Nevertheless, challenges such as limitations surrounding selectivity, reproducibility, and long-term stability still need to be resolved in order to guarantee dependable and precise detection of pesticide residues in practical, real-world scenarios [19].
To deal with the energy crisis, various technologies such as hydrogen production devices [23,24], energy storage devices (super-capacitors) [25,26], batteries [27], light emitting diodes [28], electrochromic devices [29,30], and solar cells [31] have been developed. In particular, solar cells possess unique advantages over energy-related optoelectronic devices due to their simple methods of fabrication, cost-effectiveness, and eco-friendliness [32]. Dye sensitized solar cells (DSSCs) have received extensive attention because of their stability, performance, low cost, and environmentally friendly nature [33]. DSSCs consist of various components, which can be categorized as a photoanode, light absorber material, electrolyte, and counter electrode [34]. Counter electrodes typically consist of platinum (Pt), which catalyzes the redox process during the operation of DSSCs. Specifically, counter electrode materials accept electrons from the photoanode and transfer them to the to the redox electrolyte by completing the electric circuit [35]. Pt has been proven to be the most promising and efficient counter electrode material for DSSCs and their applications [36]. Unfortunately, Pt is a precious metal and is of high cost and low abundance [37]. Thus, it is difficult to commercialize Pt-based DSSCs, and it is necessary to discover alternative Pt-free counter electrode materials for DSSCs. In this regard, various transition metal oxides such as nickel oxide [38,39], manganese dioxide [40], tin oxide [41], copper oxide [42], cobalt oxide [43], zinc oxide [44], tungsten oxide [44], and polymers such as polyaniline [45], polypyrrole [46], and polythiophene [47] have been explored as potential Pt-free counter electrode materials. Some of these Pt-free counter electrode materials have shown good photovoltaic performance, but overall, the performance of Pt-free DSSCs needs to be improved.
Recently, transition metal dichalcogenides (TMDs) such as molybdenum disulfide (MoS2), tungsten disulfide (WS2), and molybdenum diselenide (MoSe2) have attracted research interest due to their excellent optical and electrical properties [34,48,49]. MoS2 has been widely used as an electrocatalyst for the fabrication of electrochemical sensors due to its excellent catalytic properties [50]. The presence of covalent bonding between Mo and S and van der Waals bonds between the layers in this compound make it a suitable candidate for sensing applications [37]. Previously, MoS2 and its hybrid composite materials have been used in the fabrication of dopamine [50], uric acid [51], ascorbic acid [52], nitrophenol [53], hydrazine [54], and glucose [55]. The above discussed results indicated the presence of excellent electrocatalytic properties in MoS2. Moreover, MoS2 has a work function of 5.3 eV, which is comparable to Pt [56,57]. Thus, it can be assumed that MoS2 can be used as a catalyst for sensing and DSSC applications.
This paper reports the bi-functional properties of hydrothermally grown MoS2 for use in the construction of a TBZ sensor and Pt-free DSSCs. The fabricated sensor exhibited a reasonably good limit of detection, with good sensitivity for the presence of TBZ. The MoS2 based DSSCs also demonstrated good power conversion efficiency and open circuit voltage.

2. Results and Discussion

2.1. Materials Characterization

Figure 1 exhibits the P-XRD pattern of the as-obtained MoS2 sample. The P-XRD pattern of the synthesized MoS2 shows diffraction peaks at 13.89°, 32.99°, 39.46°, 49.52°, and 58.88°, which can be assigned to the presence of (002), (100), (103), (105), and (110) diffraction planes, respectively. The above peaks exist due to the interference taking place within the layers.
The surface morphology of the synthesized MoS2 was examined utilizing scanning electron microscopy (SEM). The SEM images of the as-synthesized MoS2 are displayed in Figure 2a,b. It was observed that MoS2 has sheet-like surface morphological characteristics. Thus, it can be said that MoS2 has a uniform sheet-like surface structure, which consists of flower-like self-assembled MoS2 particles. TEM images of the MoS2 were also recorded to further confirm its morphological features. Figure S1a,b shows that MoS2 has flower-like sheets, which is in agreement with the SEM results.
The phase purity and presence of the Mo and S elements in the compound were also examined using energy X-ray dispersive spectroscopy (EDX). The EDX spectrum and EDX mapping images of the synthesized MoS2 are summarized in Figure 3a–d. Figure 3b demonstrates the presence of Mo and S elements and shows that no other element has been detected. This suggests that the MoS2 sample has good phase purity, and no residual element or impurities were present in the synthesized MoS2 sample. The uniform distribution of the Mo and S elements can be seen in Figure 3c and Figure 3d, respectively. The EDX mapping results suggest uniform particle distribution. Furthermore, a high resolution XPS spectrum for Mo3d and S2p was also recorded. Figure S2a shows the obtained Mo3d XPS spectrum of the MoS2. The Mo3d spectrum shows XPS peaks at the binding energies of 231.37 eV and 228.21 eV, which can be assigned to Mo(IV)3/2 and Mo(IV)/5/2, respectively. However, the peak at 235.41 eV was ascribed to the presence of Mo(VI)3/2. Figure S2b shows the S2p XPS spectrum of MoS2, and two XPS peaks were observed at binding energies of 161.8 eV and 160.6 eV. These peaks were ascribed to the presence of S2p1/2 and S2p3/2, respectively. These results confirmed the formation of MoS2.
The specific surface area may play a vital role in the catalysis applications. Thus, it is of great importance to know the specific surface area of the as-prepared MoS2 sample. The BET method was adopted to calculate the specific surface area of the as-prepared MoS2 sample. The nitrogen (N2) adsorption-desorption isotherm of MoS2 was also recorded to obtain the specific surface area of the synthesized MoS2 sample.
The N2 adsorption-desorption isotherm of MoS2 is depicted in Figure 4. The obtained results demonstrate that MoS2 has a specific surface area of 57.5 m2/g. The aforementioned characterization studies showed that MoS2 has a layered structure with good specific surface area and sheet-like surface morphology. These properties may be beneficial for improving electron transportation. Thus, the as-synthesized MoS2 has been explored as a catalyst for the development of a TBZ sensor and Pt-free DSSCs.

2.2. Electrochemical Performance of the MSGC

The MoS2-modified GC (MSGC) electrode was used as a working electrode, with silver/silver chloride (Ag/AgCl) as a reference electrode and Pt as a counter electrode. The bare GC electrode was also used as a working electrode for the control experiment. The sensing performance of the bare GC electrode was evaluated for 20 µM TBZ in the presence of 0.1 M PBS of pH 2.0 at an applied scan rate of 50 mVs−1. Linear sweep voltammetry (LSV) was used as the sensing technology. The bare GC exhibited a current response of 8.7 µA for the oxidation of TBZ. The obtained LSV graph of the bare GC electrode has been illustrated in Figure 5a, and its oxidation peak value is presented in Figure 5b. In further studies, the LSV of the MSGC was also recorded under similar conditions and concentration as was used for the bare GC electrode. It can be noted that the MSGC exhibits an improved current response of 20.4 µA for the oxidation of 20 µM TBZ, as shown in Figure 5a,b. This interesting and significant improvement in the current response of MSGC may be attributed to the presence of electrochemically active MoS2 catalyst with high surface area on the surface of GC electrode. This suggests the presence of good catalytic properties in the hydrothermally prepared MoS2.
The concentration of TBZ present may play a significant role in the catalytic activity of the MSGC. It is of great importance to study the effect of various concentrations of TBZ on the catalytic behavior of the MSGC. Thus, we have recorded LSV graphs of the MSGC in the presence of different concentrations (0.03, 0.1, 0.3, 0.5, 0.7, 0.9, 1.1, 1.3, 1.5, 2, 3, 5, 7, 9, 11, 13, 15, and 20 µM) of TBZ. The obtained LSV graphs have been summarized in Figure 6a. The obtained results show that current response increases with increasing concentration of TBZ. The MSGC exhibits decent linear range from 0.03 to 2 µM, and 2 to 20 µM. The linear calibration graph was plotted between the oxidation current responses of the MSGC with respect to the concentration of TBZ, as shown in Figure 6b.
The charge transfer kinetic process of the MSGC was also studied for 20 µM TBZ at different applied scan rates (50–500 mV/s). Figure 7a shows the obtained LSV curves of the MSGC at different applied scan rates and reveals that current response was enhanced as the applied scan rate rose from 50 to 500 mV/s. The calibration plot between the square root of the applied scan rate and oxidation current response of the MSGC is presented in Figure 7b. It was noted that current response linearly increases with a regression coefficient (R2) value of 0.99. This suggests that the sensing of TBZ at the MSGC surface may occur via a diffusion controlled-process.
The reproducibility of the MSGC was examined for 20 µM TBZ (scan rate = 50 mV/s). The results suggested a good response, as described in Figure 8a. For the reproducibility study, five MSGC electrodes were fabricated under the same conditions and their performance was examined for the same analyte solution (20 µM TBZ, 0.1 M PBS, and pH = 2.0). The repeatability of the MSGC was also determined via running 40 consecutive LSV responses. The current values for TBZ oxidation for different cycle numbers (1, 10, 20, 30, and 40) are summarized in Figure 8b. The MSGC exhibited good repeatability for 40 cycles. The storage stability is also an important feature of electrochemical sensors, and thus the storage stability of the MSGC was studied for 28 days. The MSGC was stored for 28 days, and then its current response for the oxidation of 20 µM TBZ was examined. Observations revealed good stability at 28 days (Figure 8c).
Selectivity is one of the most important properties for any sensor in practical applications. The LSV response of the MSGC for 10 µM TBZ was recorded. Further, 50 µM of interfering compounds (AA + UA + DA + Glu + 4-CP) were introduced to the TBZ solution (Figure 9a). The obtained results showed no change in the LSV response. Similarly, a second group comprising 50 µM of alternative interfering compounds (CA + Cl + Na+ + AtCl + HZ + H2O2) were spiked to the TBZ solution and no significant change in the LSV response was observed. Thus, it was revealed that MSGC has reasonable selectivity for the detection of TBZ. The selectivity responses are shown in Figure 9a,b.
The probable mechanism for the sensing of TBZ is described in Scheme 2. In the first step, TBZ is oxidized and releases one electron during the electro-oxidation process. The positively charged TBZ further accepts the released electron to form TBZ. It can be seen that electro-oxidation of TBZ at the MSGC surface is a reversible process.
The performance of the MSGC was calculated using the equations listed below:
Limit of detection (LOD) = 3 × σe/S,
where σe is the standard error for the calibration plot between the peak current response and concentration of the TBZ, and S is the slope of the calibration plot between the peak current response and concentration of the TBZ.
Sensitivity = S/area of the GC electrode
The MSGC demonstrated good LOD of 0.1 µM and sensitivity of 7.47 µA/µM.cm2. The obtained results have been summarized in Table 1, and are reasonable compared to those in the literature [19,58,59,60].

2.3. DSSC Applications

In the first step, we examined the catalytic behavior of FTO@MoS2 in the presence of liquid electrolyte solution (0.05 M LiI, 0.01 M I2, and 0.5 M LiClO4 in acetonitrile; scan rate = 50 mV/s). The obtained CV curve of FTO@MoS2 is displayed in Figure 10a. The obtained results exhibit good electrocatalytic activity towards the redox reaction at the surface of FTO@MoS2. The CV of FTO@Pt was also recorded under similar conditions as were used for FTO@MoS2. FTO@Pt exhibited higher electrocatalytic redox properties compared to FTO@MoS2. However, the electrocatalytic activity of FTO@MoS2 is overall reasonable compared to the FTO@Pt.
Thus, it can be concluded that FTO@MoS2 can be applied as a counter electrode for the fabrication of DSSCs. A schematic diagram of the DSSC is shown in Figure 10b. The redox reactions at the surface of FTO@MoS2 can be seen in the inset of Figure 10b. The photovoltaic performance of DSSCs can be evaluated in terms of power conversion efficiency (PCE). The photocurrent density versus voltage (JV) results of the FTO@Pt and FTO@MoS2 counter electrode-based DSSCs were recorded under 1 sun conditions. The obtained JV results are shown in Figure 11. The photovoltaic parameters of the FTO@Pt and FTO@MoS2 counter electrode-based DSSCs are summarized in Figure 12a–d.
It can be seen that FTO@MoS2 counter electrode-based DSSCs exhibited open circuit voltage (Voc) of 0.73 V (Figure 12a) with photocurrent density (Jsc) of 13.48 mA/cm2 (Figure 12c). The fill factor (FF) of the photovoltaic cells measures the quality of the cell. The FF has significant impact on the efficiency of DSSCs. The FF can be defined as given below:
F F = P m a x V o c × I s c ,
where Pmax is the maximum power point and Isc is the short circuit current.
The FTO@Pt counter electrode-based DSSCs exhibit an FF of 0.52, whereas FTO@MoS2 counter electrode-based DSSCs exhibit an FF of 0.51 (Figure 12b). A Jsc of 14.82 mA/cm2 was obtained for FTO@Pt counter electrode-based DSSCs, as shown in Figure 12c. A PCE of 5.8% and 5.01% was obtained for FTO@Pt and FTO@MoS2 counter electrode-based DSSCs, respectively (Figure 12d).

3. Materials and Methods

3.1. Materials

Sodium molybdate dihydrate (≥99.5%) was purchased from Merck. Polyvinylidene fluoride (PVDF) was bought from Alfa Aesar, India. Thiourea (ACS reagent, ≥99.0%), Nafion (5 wt%), fluorine doped tin oxide (FTO) glass, N3 (cis-bis(isothiocyanato)bis(2,2′-bipyridyl-4,4′-dicarboxylato)ruthenium(II)) dye (95%), and chloroplatinic acid hexahydrate (H2[Pt(Cl)6]·6H2O) (ACS reagent, ≥37.50% Pt basis) were purchased from Merck. Titanium tetrachloride (TiCl4, ReagentPlus®, 99.9% trace metals basis) in 1 molar toluene (anhydrous, 99.8%), polyethylene glycol (PEG, molecular weight = 10,000), iodine (I2, ACS reagent, ≥99.8%), triton-X (X-100), isopropyl alcohol (IPA), lithium iodide (LiI, 99.9% trace metals basis), acetylacetone (CH3COCH2COCH3), and thiabendazole (TBZ, ≥99%) were purchased from Sigma, India. Titanium dioxide (TiO2, 99%) nanopowder (P25 Deguusa) was purchased from Nanoshel, India. Other used chemicals and reagents including phosphate buffer solutions (PBS) were purchased from Sigma and Alfa Aesar.

3.2. Hydrothermal Synthesis of MoS2

The present work adopted the hydrothermal method for the fabrication of MoS2, which has been reported elsewhere [28]. In brief, 1.6 g thiourea was dissolved in 25 mL deionized water (DI water). In another beaker, 1.4 g of sodium molybdate was dissolved in 25 mL DI water. The aqueous solution of thiourea was added to the aqueous solution of sodium molybdate. This reaction mixture was stirred at room temperature for 30 min, transferred to a Teflon cup, and covered tightly with a stainless steel autoclave. The autoclave was put into the muffle furnace and heated at 180 °C for 24 h. The muffle furnace was cooled down to room temperature following this. After cooling down the muffle furnace, the autoclave was opened. MoS2 was collected using centrifugation and washed with DI water and ethanol. After washing, the MoS2 was dried in a vacuum oven at 60 °C for 12 h. The synthesis of MoS2 is schematically described in Scheme 1.

3.3. Fabrication of MSGC

The synthesized MoS2 (1.5 mgmL−1) was dispersed in DI water via ultrasonication for 2 h (0.1 wt% of the Nafion was added as a binder). 7 µL of the prepared dispersion was drop-casted on the active surface area (3 mm) of the glassy carbon (GC) electrode and dried at RT for 3 h. The modified GC electrode was used as the working electrode, and Ag/AgCl was adopted as the reference electrode. A Pt wire-based electrode was used as a counter electrode. The sensing activity of the MoS2 modified GC electrode (MSGC) was determined in a three electrode assembly. A CH instruments-based potentiostat system was used as the sensing platform. The fabrication of MSGC is described in Scheme 2.
A selectivity test was performed using 10 µM TBZ in the presence of 50 µM of interfering compounds. There were two groups of interfering compounds tested. Group A contained interfering molecules such as ascorbic acid (AA, 10 µM), uric acid (UA, 10 µM), dopamine (DA, 10 µM), glucose (Glu, 10 µM), and 4-chlorophenol (4-CP, 10 µM). In group B, 50 µM of interfering compounds such as citric acid (CA, 10 µM), chloride (Cl, 5 µM), sodium (Na+,5 µM), acetylthiocholine chloride (AtCl, 10 µM), hydrazine (HZ, 10 µM), and hydrogen peroxide (H2O2, 10 µM) were used.

3.4. Fabrication of DSSCs

In the first step, a photoanode was fabricated. FTO coated glass substrates (Sigma, 2 × 2 cm pieces) were cleaned in an ultrasonic bath using a mild alkaline solution. The cleaning of the FTO substrates was completed via the use of DI water, acetone, and isopropyl alcohol for 10 min each. The blocking layer of TiO2 was coated on to the cleaned FTO to stop recombination reactions. The FTO glass substrate was immersed (facing conductive side upwards) in a 45 mM aqueous solution of TiCl4 for 25 min at 60 °C. The TiCl4-immersed FTO glass substrate was heated at 100 °C for 10 min. The doctor-blade method was used for the fabrication of the TiO2 layer, using an invisible tape as the spacer to maintain the uniform thickness of the TiO2 film. The TiO2 paste was prepared by mixing 0.15 mg of polyethylene glycol (PEG; MW 10,000) in 1 molar HNO3, followed by a spike of 0.21 mL acetylacetone, 0.04 mL Triton-X, and TiO2 (P25 Deguusa) powder. This solution was sonicated for 60 min and stirred for 1 day to obtain the colloidal TiO2 solution. The TiO2 coated FTO substrates were sintered at 450 °C for 40 min. In another step, the counter electrode was fabricated using MoS2 as the base material and Pt as a control counter electrode. The MoS2 (15 mg) was dispersed in 2-propanol for 30 min using sonication. 50 µL of the prepared dispersion was drop-casted on to the pre-cleaned FTO glass substrate, which was used as a Pt-free counter electrode. The Pt counter electrode was also fabricated as a control. A 5 mM solution of H2[Pt(Cl)6].6H2O was prepared in 2-propanol, drop-casted on to the pre-cleaned FTO glass substrates, and sintered at 450 °C for 30 min.
The fabricated photoanode (FTO/TiO2) was dipped in the N3 dye solution [5 mM N3 dye in acetonitrile/tert-butanol (ratio = 1:1)] for 1 day under dark conditions. After 24 h of dipping in N3 dye solution, the photoanode was rinsed with ethanol to remove the extra dye molecules from the photoanode surface. In the final step, the photoanode and counter electrode were clipped together to assemble the DSSCs. Parafilm was used as a separator to avoid contact/short circuit between the two conductive surfaces. The liquid electrolyte was prepared using 0.05 M I2, 0.5 M LiI, and 0.5 M 4-tert-butylpyridine in 3-methoxypropionitrile. The liquid electrolyte was injected in between the photoanode and counter electrode.

3.5. Instruments

The XRD pattern of the prepared MoS2 was recorded on a Rigaku instrument (Cu/Kα radiation (wavelength = 1.54 Å)) (Japan). Scanning electron microscope (SEM) images were captured on a Zeiss Supra 55 microscope. Energy dispersive X-ray (EDX) studies were carried out using an Oxford EDX instrument. The Brunauer-Emmett-Teller (BET) analysis was carried out using nitrogen isotherm (adsorption/desorption) on Quantachrome Instruments: Autosorb iQ (version 1.11). The cyclic voltammograms (CV) of the counter electrodes were obtained in the presence of liquid electrolyte solution (0.05 M LiI, 0.01 M I2, and 0.5 M LiClO4 in acetonitrile; scan rate = 50 mV/s). The efficiency of the fabricated DSSCs was examined using a solar simulator (Xe arc lamp, 100 mW/cm2, AM 1.5 G). The XPS scan was obtained using a Thermo Fischer Scientific Xray photoelectron spectrometer. The transmission electron microscope images were obtained using a TEM CM 200 instrument.

4. Conclusions

Finally, it can be concluded that MoS2 flowers have been synthesized using the hydrothermal method and characterized using various sophisticated techniques. The synthesized MoS2 has a high surface area and sheet-like surface morphology. Further, an electrochemical sensor was fabricated using MoS2 as an electrocatalyst for the determination of the presence of TBZ. The MoS2-modified GC electrode exhibited a limit of detection (LOD) of 0.1 µM. The MoS2-modified GC electrode also showed good selectivity and sensitivity for the determination of the presence of TBZ. MoS2 counter electrode-based dye sensitized solar cells were also fabricated. The fabricated dye sensitized solar cells showed good efficiency of 5.01%. This work reported the bi-functional applications of hydrothermally synthesized MoS2 for the fabrication of TBZ sensors and Pt-free dye sensitized solar cells.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14020107/s1: Figure S1. TEM images (a,b) of the MoS2 and Figure S2. Mo3d (a) and S2p (b) XPS scan of MoS2.

Author Contributions

Conceptualization, M.Q.K. and W.R.; methodology, K.A. and R.A.K.; investigation, K.A. and R.A.K.; resources, A.P.; writing—original draft, M.S., A.P. and R.A.K.; writing—review and editing, R.A.K. and M.S.; funding acquisition, K.A., R.A.K. and M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by project number RSP2024R400, King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

Data are contained within the article.

Acknowledgments

R.A.K. gratefully acknowledge the researchers supporting project (project number RSP2024R400), King Saud University, Riyadh, Saudi Arabia. A.P. is grateful to FCT and IST, Portugal, for financial support through “DL/57/2017” (contract no. IST-ID/197/2019, DOI 10.54499/DL57/2016/CP1384/CT0081). This work was supported by the Fundação para a Ciência e a Tecnologia (FCT), Portugal, in the form of projects UIDB/00100/2020 and UIDP/00100/2020 of Centro de Química Estrutural. M.S. is grateful to Faculdade de Engenharia, Universidade Lusófona—Centro Universitário de Lisboa for providing laboratory facilities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Mohammad, A.; Ansari, S.N.; Chaudhary, A.; Ahmad, K.; Rajak, R.; Tauqeer, M.; Mobin, S.M. Enthralling Adsorption of Different Dye and Metal Contaminants from Aqueous Systems by Cobalt/Cobalt Oxide Nanocomposites Derived from Single-Source Molecular Precursors. Chem. Sel. 2018, 3, 5733–5741. [Google Scholar] [CrossRef]
  2. Raza, W.; Ahmad, K.; Kim, H. Nitrogen-doped graphene as an efficient metal-free catalyst for ammonia and non-enzymatic glucose sensing. J. Phys. Chem. Solids 2022, 160, 110359. [Google Scholar] [CrossRef]
  3. Khan, M.Q.; Ahmad, K.; Alsalme, A.; Kim, H. Hydrothermal synthesis of nanostructured NiO for hydrazine sensing application. Mater. Chem. Phys. 2022, 289, 126463. [Google Scholar] [CrossRef]
  4. He, X.; Hwang, H.-M. Nanotechnology in food science: Functionality. applicability, and safety assessment. J. Food Drug Anal. 2016, 24, 671–681. [Google Scholar] [CrossRef]
  5. Zhao, J.; Dang, Z.; Muddassir, M.; Raza, S.; Zhong, A.; Wang, X.; Jin, J. A New Cd(II)-Based Coordination Polymer for Efficient Photocatalytic Removal of Organic Dyes. Molecules 2023, 28, 6848. [Google Scholar] [CrossRef]
  6. Zheng, X.; Guo, L.; Zhu, C.; Hu, T.; Gong, X.; Wu, C.; Wang, G.; Dong, H.; Hou, Y. A robust electrochemical sensor based on AgNWs@MoS2 for highly sensitive detection of thiabendazole residues in food samples. Food Chem. 2024, 433, 137304. [Google Scholar] [CrossRef]
  7. Lengai, G.M.W.; Muthomi, J.W.; Mbega, E.R. Phytochemical activity and role of botanical pesticides in pest management for sustainable agricultural crop production. Sci. Afr. 2020, 7, e00239. [Google Scholar] [CrossRef]
  8. Zhong, W.; Zou, J.; Yu, Q.; Gao, Y.; Qu, F.; Liu, S.; Zhou, H.; Lu, L. Ultrasensitive indirect electrochemical sensing of thiabendazole in fruit and water by the anodic stripping voltammetry of Cu2+ with hierarchical Ti3C2Tx−TiO2 for signal amplification. Food Chem. 2023, 402, 134379. [Google Scholar] [CrossRef]
  9. Wang, H.; Cai, L.; Hao, W.; Wang, Y.; Fang, G.; Wang, S. Melamine-Ag with dual functions of electrochemiluminescence luminophore and coreactant accelerators: Construction of MIP/M-Ag@MoS2-QDs sensing platform for specific detection of thiabendazole. Food Chem. 2023, 425, 136472. [Google Scholar] [CrossRef]
  10. Kazemifard, N.; Ensafi, A.A.; Rezaei, B. Green synthesized carbon dots embedded in silica molecularly imprinted polymers, characterization and application as a rapid and selective fluorimetric sensor for determination of thiabendazole in juices. Food Chem. 2020, 310, 125812. [Google Scholar] [CrossRef] [PubMed]
  11. Zheng, D.; Hu, X.; Fu, X.; Xia, Z.; Zhou, Y.; Peng, L.; Yu, Q.; Peng, X. Flowerlike Ni–NiO composite as magnetic solid-phase extraction sorbent for analysis of carbendazim and thiabendazole in edible vegetable oils by liquid chromatography-mass spectrometry. Food Chem. 2022, 374, 131761. [Google Scholar] [CrossRef]
  12. Peng, X.-X.; Bao, G.-M.; Zhong, Y.-F.; Zhang, L.; Zeng, K.-B.; He, J.-X.; Xiao, W.; Xia, Y.-F.; Fan, Q.; Yuan, H.-Q. Highly sensitive and rapid detection of thiabendazole residues in oranges based on a luminescent Tb3+−functionalized MOF. Food Chem. 2021, 343, 128504. [Google Scholar] [CrossRef]
  13. Słowik-Borowiec, M.; Szpyrka, E.; Walorczyk, S. Gas chromatographic determination of pesticide residues in white mustard. Food Chem. 2015, 173, 997–1005. [Google Scholar] [CrossRef]
  14. Piendl, S.K.; Geissler, D.; Weigelt, L.; Belder, D. Multiple heart-cutting twodimensional chip-HPLC combined with deep-UV fluorescence and mass spectrometric detection. Anal. Chem. 2020, 92, 3795–3803. [Google Scholar] [CrossRef]
  15. Zhao, Y.; Deng, G.; Liu, X.; Sun, L.; Li, H.; Cheng, Q.; Xi, K.; Xu, D. MoS2/Ag nanohybrid: A novel matrix with synergistic effect for small molecule drugs analysis by negativeion matrix-assisted laser desorption/ionization time-of-flight mass spectrometry. Anal. Chim. Acta 2016, 937, 87–95. [Google Scholar] [CrossRef]
  16. Gough, K.C.; Patel, S.; Baker, C.A.; Maddison, B.C. Development of immunoassays for the detection of the fungicide penconazole and its urinary metabolite. J. Agricult. Food Chem. 2009, 57, 9393–9399. [Google Scholar] [CrossRef]
  17. Ahmad, K.; Raza, W.; Alsulmi, A.; Kim, H. Fabrication of electrochemical sensor for metronidazole using MoS2/graphite-like carbon nitride composite modified glassy carbon electrode. Diamond Relat. Mater. 2023, 138, 110178. [Google Scholar] [CrossRef]
  18. Yang, J.; Zhang, D.; Wang, L.; Long, N.; Zhang, M.; Zhang, L. An electrochemical method for high sensitive detection of thiabendazole and its interaction with human serum albumin. Food Anal. Methods 2015, 8, 507–514. [Google Scholar] [CrossRef]
  19. Zhang, B.; Chen, Q.; Liu, D.; Fang, F.; Mu, M.; Xie, Y. Heterogeneous sensitization from nanoporous gold and titanium carbide (MXene) combining with molecularly imprinted polymers for highly sensitive and specific sensing detection of thiabendazole. Sens. Actuators B Chem. 2022, 367, 132159. [Google Scholar] [CrossRef]
  20. Khan, M.Q.; Khan, R.A.; Alsalme, A.; Ahmad, K.; Kim, H. Design and Fabrication of α-MnO2-Nanorods-Modified Glassy-Carbon-Electrode-Based Serotonin Sensor. Biosensors 2022, 12, 849. [Google Scholar] [CrossRef]
  21. Khan, M.Q.; Kumar, P.; Khan, R.A.; Ahmad, K.; Kim, H. Fabrication of Sulfur-Doped Reduced Graphene Oxide Modified Glassy Carbon Electrode (S@rGO/GCE) Based Acetaminophen Sensor. Inorganics 2022, 10, 218. [Google Scholar] [CrossRef]
  22. Kumar, P.; Khan, M.Q.; Khan, R.A.; Ahmad, K.; Kim, H. Hydrothermal Synthesis of MnO2/Reduced Graphene Oxide Composite for 4-Nitrophenol Sensing Applications. Inorganics 2022, 10, 219. [Google Scholar] [CrossRef]
  23. Ye, D.; Liu, L.; Peng, Q.; Qiu, J.; Gong, H.; Zhong, A.; Liu, S. Effect of Controlling Thiophene Rings on D-A Polymer Photocatalysts Accessed via Direct Arylation for Hydrogen Production. Molecules 2023, 28, 4507. [Google Scholar] [CrossRef]
  24. Huang, X.; Chen, N.; Ye, D.; Zhong, A.; Liu, H.; Li, Z.; Liu, S. Structurally complementary star-shaped unfused ring electron acceptors with simultaneously enhanced device parameters for ternary organic solar cells. Sol. RRL 2023, 7, 2300143. [Google Scholar] [CrossRef]
  25. Behzadi Pour, G.; Fekri Aval, L.; Kianfar, E. Comparative studies of nanosheet-based supercapacitors: A review of advances in electrodes materials, Case Studies Chem. Environ. Eng. 2024, 9, 100584. [Google Scholar]
  26. Gong, K.; Lee, H.; Choi, Y.; Jung, G.; Keum, K.; Kim, J.W.; Ha, J.S. A flexible supercapacitor with high energy density and wide range of temperature tolerance using a high-concentration aqueous gel electrolyte. Electrochim. Acta 2024, 475, 143585. [Google Scholar] [CrossRef]
  27. Chen, S.; Cai, Z.; Wang, R.; Li, T.; Song, J.; Yuan, L. Ultrathin transparent carbon nanotube/LiMn2O4 and carbon nanotube/MoS2 film electrodes for Li-ion full batteries. Solid State Ion. 2023, 399, 116320. [Google Scholar] [CrossRef]
  28. Lai, M.T.L.; Lee, K.M.; Yang, T.C.K.; Lai, C.W.; Chen, C.-Y.; Johan, M.R.; Juan, J.C. Highly effective interlayer expanded MoS2 coupled with Bi2WO6 as p-n heterojunction photocatalyst for photodegradation of organic dye under LED white light. J. Alloys Compd. 2023, 953, 169834. [Google Scholar] [CrossRef]
  29. Ahmad, K.; Shinde, M.A.; Song, G.; Kim, H. Fabrication of MoSe2/AgNWs/PET electrode for flexible electrochromic smart window applications. Opt. Mater. 2022, 132, 112805. [Google Scholar] [CrossRef]
  30. Ahmad, K.; Shinde, M.A.; Song, G.; Kim, H. Preparation of CoS2/WO3 thin films for the construction of electrochromic devices. Ceram. Int. 2023, 49, 10119–10128. [Google Scholar] [CrossRef]
  31. Karthigaimuthu, D.; Maram, P.S.; Kumar, B.A.; Ramalingam, G.; Elangovan, T.; Sangaraju, S. Hydrothermal synthesis of MoS2-Mg(OH)2-BiVO4 ternary hierarchical heterostructures for dye-sensitized solar cell application. Mater. Lett. 2024, 359, 135890. [Google Scholar] [CrossRef]
  32. Vibavakumar, S.; Nisha, K.D.; Harish, S.; Archana, J.; Navaneethan, M. Synergistic effect of MoO3/MoS2 in improving the electrocatalytic performance of counter electrode for enhanced efficiency in dye-sensitized solar cells. Mater. Sci. Semicond. Process. 2023, 161, 107431. [Google Scholar] [CrossRef]
  33. Ahmad, K.; Kim, H. MoSe2−nanoflowers as bi-functional electro-catalyst: An efficient counter electrode material for dye-sensitized solar cells and electrode modifier for hydrazine sensor. Mater. Chem. Phys. 2023, 296, 127260. [Google Scholar] [CrossRef]
  34. Xu, X.; Yang, W.; Yang, F.; Hou, L.; Li, Y. Highly active and reflective MoS2 counter electrode for enhancement of photovoltaic efficiency of dye sensitized solar cells. Electrochim. Acta 2016, 212, 614–620. [Google Scholar]
  35. Zheng, F.; Huang, N.; Peng, R.; Ding, Y.; Li, G.; Xia, Z.; Sun, P.; Sun, X.; Geng, J. Cobalt doped molybdenum disulfide in-situ grown on graphite paper with excellent electrocatalytic activity for triiodide evolution. Electrochim. Acta 2018, 263, 328–337. [Google Scholar] [CrossRef]
  36. Mahato, S.; Nandigana, P.; Pradhan, B.; Subramanian, B.; Panda, S.K. Enhanced efficiency of DSSC by lyophilized tin-doped molybdenum sulfide as counter electrode. J. Alloys Compd. 2022, 894, 162406. [Google Scholar] [CrossRef]
  37. Wang, Y.; Guo, Y.; Chen, W.; Luo, Q.; Lu, W.; Xu, P.; Chen, D.; Yin, X.; He, M. Sulfur doped reduced graphene oxide/MoS2 composite with exposed active sites as efficient Pt-free counter electrode for dye-sensitized solar cell. Appl. Surf. Sci. 2018, 452, 232–238. [Google Scholar] [CrossRef]
  38. Silambarasan, K.; Archana, J.; Athithya, S.; Harish, S.; Ganesh, R.S.; Navaneethan, M.; Ponnusamy, S.; Muthamizhchelvan, C.; Hara, K.; Hayakawa, Y. Hierarchical NiO@NiS@graphene nanocomposite as a sustainable counter electrode for Pt free dye-sensitized solar cell. Appl. Surf. Sci. 2020, 501, 144010. [Google Scholar] [CrossRef]
  39. Sarkar, A.; Chakraborty, A.K.; Bera, S. NiS/rGO nanohybrid: An excellent counter electrode for dye sensitized solar cell. Sol. Energy Mater. Sol. Cells 2018, 182, 314–320. [Google Scholar] [CrossRef]
  40. Aljafari, B.; Vijaya, S.; Takshi, A.; Anandan, S. Copper doped manganese dioxide as counter electrode for dye-sensitized solar cells. Arab. J. Chem. 2022, 15, 104068. [Google Scholar] [CrossRef]
  41. Cui, X.; Xu, W.; Xie, Z.; Wang, Y. Hierarchical SnO2@SnS2 Counter Electrodes for Remarkable High-efficiency Dye-sensitized Solar Cells. Electrochim. Acta 2015, 186, 125–132. [Google Scholar] [CrossRef]
  42. Alami, A.H.; Rajab, B.; Abed, J.; Faraj, M.; Hawili, A.A.; Alawadhi, H. Investigating various copper oxides-based counter electrodes for dye sensitized solar cell applications. Energy 2019, 174, 526–533. [Google Scholar] [CrossRef]
  43. Chen, L.; Chen, W.; Wang, E. Graphene with cobalt oxide and tungsten carbide as a low-cost counter electrode catalyst applied in Pt-free dye-sensitized solar cells. J. Power Sources 2018, 380, 18–25. [Google Scholar] [CrossRef]
  44. Mahajan, P.; Datt, R.; Gupta, V.; Arya, S. Synthesis and characterization of ZnO@WO3 core/shell nanoparticles as counter electrode for dye-sensitized solar cell. Surf. Interfaces 2022, 30, 101920. [Google Scholar] [CrossRef]
  45. Ravichandran, S.; Varthamanan, Y.; Akilandeswari; Elangoven, T.; Ragupathi, C.; Murugesan, S. Effect of polyaniline/FeS2 composite and usages of alternates counter electrode for dye-sensitized solar cells. Mater. Today Proceed. 2022, 49, 2615–2619. [Google Scholar] [CrossRef]
  46. Ahmed, U.; Shahid, M.M.; Shahabuddin, S.; Rahim, N.A.; Alizadeh, M.; Pandey, A.K.; Sagadevan, S. An efficient platform based on strontium titanate nanocubes interleaved polypyrrole nanohybrid as counter electrode for dye-sensitized solar cell. J. Alloys Compd. 2021, 860, 158228. [Google Scholar] [CrossRef]
  47. Asok, A.; Haribabu, K. Synthesis and performance of polythiophene-iridium oxide composite as counter electrode in dye sensitized solar cell. Curr. Appl. Phys. 2023, 49, 64–69. [Google Scholar] [CrossRef]
  48. Althomali, R.H.; Al-Hawary, S.I.S.; Gehlot, A.; Qasim, M.T.; Abdullaeva, B.; Sapaev, I.B.; Al-Kharsan, I.H.; Alsalamy, A. A novel Pt-free counter electrode based on MoSe2 for cost effective dye-sensitized solar cells (DSSCs): Effect of Ni doping. J. Phys. Chem. Solids 2023, 182, 111597. [Google Scholar] [CrossRef]
  49. Ahmad, M.S.; Rahim, N.A.; Shahabuddin, S.; Mehmood, S.; Khan, A.D. Effect of WS2 nano-sheets on the catalytic activity of polyaniline nano-rods based counter electrode for dye sensitized solar cell. Phys. E Low-Dimens. Syst. Nanostruct. 2021, 126, 114466. [Google Scholar] [CrossRef]
  50. Sabar, M.; Amara, U.; Riaz, S.; Hayat, A.; Nasir, M.; Nawaz, M.H. Fabrication of MoS2 enwrapped carbon cloth as electrochemical probe for non-enzymatic detection of dopamine. Mater. Lett. 2022, 308, 131233. [Google Scholar] [CrossRef]
  51. Sha, R.; Vishnu, N.; Badhulika, S. MoS2 based ultra-low-cost, flexible, non-enzymatic and non-invasive electrochemical sensor for highly selective detection of Uric acid in human urine samples. Sens. Actuators B Chem. 2019, 279, 53–60. [Google Scholar] [CrossRef]
  52. Huang, M.; Tian, H.; Zhou, P.; Ma, S.; Chen, J.; Zhang, N.; Zhang, K. Electrochemical Sensor for Detection of Ascorbic Acid based on MoS2-AuNPs Modified Glassy Carbon Electrode. Int. Electrochem. Sci. 2021, 16, 151014. [Google Scholar] [CrossRef]
  53. Kalia, H.; Kumar, R.; Sharma, R.; Kumar, S.; Singh, D.; Singh, R.K. Two-dimensional layered rGO-MoS2 heterostructures decorated with Fe3O4 nanoparticles as an electrochemical sensor for detection of para-nitrophenol. J. Phys. Chem. Solids 2024, 184, 111719. [Google Scholar] [CrossRef]
  54. Villa-Manso, A.M.; Revenga-Parra, M.; Vera-Hidalgo, M.; Sulleiro, M.V.; Pérez, E.M.; Lorenzo, E.; Pariente, F. 2D MoS2 nanosheets and hematein complexes deposited on screen-printed graphene electrodes as an efficient electrocatalytic sensor for detecting hydrazine. Sens. Actuators B Chem. 2021, 345, 130385. [Google Scholar] [CrossRef]
  55. Sharma, K.P.; Shin, M.; Awasthi, G.P.; Cho, S.; Yu, C. One-step hydrothermal synthesis of CuS/MoS2 composite for use as an electrochemical non-enzymatic glucose sensor. Heliyon 2023, 10, e23721. [Google Scholar] [CrossRef]
  56. Kim, J.H.; Lee, J.; Kim, J.H.; Hwang, C.C.; Lee, C.; Park, J.Y. Work function variation of MoS2 atomic layers grown with chemical vapor deposition: The effects of thickness and the adsorption of water/oxygen molecules. Appl. Phys. Lett. 2015, 106, 10–15. [Google Scholar] [CrossRef]
  57. Mithari, P.A.; Mendhe, A.C.; Karade, S.S.; Sankapal, B.R.; Patrikar, S.R. MoS2 nanoflakes anchored MWCNTs: Counter electrode in dye-sensitized solar cell. Inorg. Chem. Commun. 2021, 132, 108827. [Google Scholar] [CrossRef]
  58. Ribeiro, F.W.P.; de Oliveira, R.C.; de Oliveira, A.G.; Nascimento, R.F.; Becker, H.; de Lima-Neto, P.; Correia, A.N. Electrochemical sensing of thiabendazole in complex samples using boron-doped diamond electrode. J. Electroanal. Chem. 2020, 866, 114179. [Google Scholar] [CrossRef]
  59. Mola, A.M.; Debebe, S.E.; Mekonnen, M.L. Sensitive Electrochemical Detection of Thiabendazole in Fruits Using Ag−MoS2 Electrode. Electroanalysis 2023, 35, e202200045. [Google Scholar] [CrossRef]
  60. Li, J.; Ma, X.; Zhang, M.; Li, D.; Yuan, Y.; Fan, Y.; Xie, X.; Guo, L.; Zeng, G. Preparation of Molecularly Imprinted Polymer Sensor on Electrochemically Reduced Graphene Oxide Modified Electrode for Selective Probing of Thiabendazole. J. Electrochem. Soc. 2019, 166, B84. [Google Scholar] [CrossRef]
Figure 1. P-XRD pattern of the synthesized MoS2.
Figure 1. P-XRD pattern of the synthesized MoS2.
Catalysts 14 00107 g001
Figure 2. SEM (a,b) images of the MoS2.
Figure 2. SEM (a,b) images of the MoS2.
Catalysts 14 00107 g002
Figure 3. EDX electron image (a), EDX spectrum (b) and EDX mapping image of Mo (c) and S (d) elements.
Figure 3. EDX electron image (a), EDX spectrum (b) and EDX mapping image of Mo (c) and S (d) elements.
Catalysts 14 00107 g003
Figure 4. Nitrogen adsorption-desorption isotherm of MoS2.
Figure 4. Nitrogen adsorption-desorption isotherm of MoS2.
Catalysts 14 00107 g004
Figure 5. LSV graph (a) and current value (b) of the GC and MSGC for 20 µM TBZ (conditions; 50 mVs−1; 0.1 M PBS; pH = 2.0).
Figure 5. LSV graph (a) and current value (b) of the GC and MSGC for 20 µM TBZ (conditions; 50 mVs−1; 0.1 M PBS; pH = 2.0).
Catalysts 14 00107 g005
Figure 6. LSV graphs (a) of the MSGC for various concentrations (0.03, 0.1, 0.3, 0.5, 0.7, 0.9, 1.1, 1.3, 1.5, 2, 3, 5, 7, 9, 11, 13, 15, and 20 µM) of TBZ (scan rate = 50 mV/s). Corresponding calibrations curve (b).
Figure 6. LSV graphs (a) of the MSGC for various concentrations (0.03, 0.1, 0.3, 0.5, 0.7, 0.9, 1.1, 1.3, 1.5, 2, 3, 5, 7, 9, 11, 13, 15, and 20 µM) of TBZ (scan rate = 50 mV/s). Corresponding calibrations curve (b).
Catalysts 14 00107 g006
Figure 7. LSV graphs (a) of the MSGC for 20 µM TBZ (0.1 M PBS, pH = 2.0) at different scan rates (50–500 mV/s). Corresponding calibrations curve (b).
Figure 7. LSV graphs (a) of the MSGC for 20 µM TBZ (0.1 M PBS, pH = 2.0) at different scan rates (50–500 mV/s). Corresponding calibrations curve (b).
Catalysts 14 00107 g007
Figure 8. Reproducibility (a), repeatability (b) and stability (c) of MSGC for 20 µM TBZ (scan rate = 50 mV/s).
Figure 8. Reproducibility (a), repeatability (b) and stability (c) of MSGC for 20 µM TBZ (scan rate = 50 mV/s).
Catalysts 14 00107 g008
Figure 9. LSV response (a) and peak current value (b) of MSGC for 10 µM TBZ (blue curve), 10 µM TBZ + 50 µM interfering compounds group A (AA + UA + DA+ Glu + 4-CP (black curve) and 10 µM TBZ + 50 µM interfering compounds group B (CA + Cl + Na+ + AtCl + HZ + H2O2 (red curve) at scan rate = 50 mV/s.
Figure 9. LSV response (a) and peak current value (b) of MSGC for 10 µM TBZ (blue curve), 10 µM TBZ + 50 µM interfering compounds group A (AA + UA + DA+ Glu + 4-CP (black curve) and 10 µM TBZ + 50 µM interfering compounds group B (CA + Cl + Na+ + AtCl + HZ + H2O2 (red curve) at scan rate = 50 mV/s.
Catalysts 14 00107 g009
Figure 10. CV graphs (a) of the FTO@Pt and FTO@MoS2 in presence of liquid electrolyte solution (0.05 M LiI, 0.01 M I2, and 0.5 M LiClO4 in acetonitrile; scan rate = 50 mV/s). Schematic picture (b) of the DSSCs.
Figure 10. CV graphs (a) of the FTO@Pt and FTO@MoS2 in presence of liquid electrolyte solution (0.05 M LiI, 0.01 M I2, and 0.5 M LiClO4 in acetonitrile; scan rate = 50 mV/s). Schematic picture (b) of the DSSCs.
Catalysts 14 00107 g010
Figure 11. J-V graphs of the FTO@Pt and FTO@MoS2 counter electrode based fabricated DSSCs under 1 sun conditions (AM 1.5 G).
Figure 11. J-V graphs of the FTO@Pt and FTO@MoS2 counter electrode based fabricated DSSCs under 1 sun conditions (AM 1.5 G).
Catalysts 14 00107 g011
Figure 12. Photovoltaic parameters: Voc (a), FF (b), Jsc (c) and PCE (d) of the FTO@Pt and FTO@MoS2 counter electrode-based DSSCs.
Figure 12. Photovoltaic parameters: Voc (a), FF (b), Jsc (c) and PCE (d) of the FTO@Pt and FTO@MoS2 counter electrode-based DSSCs.
Catalysts 14 00107 g012
Scheme 1. Schematic representation of the synthesis of MoS2.
Scheme 1. Schematic representation of the synthesis of MoS2.
Catalysts 14 00107 sch001
Scheme 2. Schematic illustration of the fabrication of MSGC for application in a TBZ sensor.
Scheme 2. Schematic illustration of the fabrication of MSGC for application in a TBZ sensor.
Catalysts 14 00107 sch002
Table 1. Comparison of the LOD and linear range of the MSGC with previously reported TBZ sensors.
Table 1. Comparison of the LOD and linear range of the MSGC with previously reported TBZ sensors.
Sensing MaterialLOD (µM)Linear Range (µM)Sensitivity (µA/µM.cm2)References
MSGC0.10.03 to 2 and 2 to 207.47This study
Au@Ti3C2Tx0.231–80-[19]
Boron-diamond doped electrode0.120.49–11.2-[58]
Ag−MoS20.11–10-[59]
Molecularly imprinted polymer/reduced graphene oxide0.1250.5–10, and 10–120-[60]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khan, M.Q.; Ahmad, K.; Raza, W.; Khan, R.A.; Sutradhar, M.; Paul, A. Hydrothermally Synthesized MoS2 as Electrochemical Catalyst for the Fabrication of Thiabendazole Electrochemical Sensor and Dye Sensitized Solar Cells. Catalysts 2024, 14, 107. https://doi.org/10.3390/catal14020107

AMA Style

Khan MQ, Ahmad K, Raza W, Khan RA, Sutradhar M, Paul A. Hydrothermally Synthesized MoS2 as Electrochemical Catalyst for the Fabrication of Thiabendazole Electrochemical Sensor and Dye Sensitized Solar Cells. Catalysts. 2024; 14(2):107. https://doi.org/10.3390/catal14020107

Chicago/Turabian Style

Khan, Mohd Quasim, Khursheed Ahmad, Waseem Raza, Rais Ahmad Khan, Manas Sutradhar, and Anup Paul. 2024. "Hydrothermally Synthesized MoS2 as Electrochemical Catalyst for the Fabrication of Thiabendazole Electrochemical Sensor and Dye Sensitized Solar Cells" Catalysts 14, no. 2: 107. https://doi.org/10.3390/catal14020107

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop