Uneingeschränkter Zugang

A Highly Selective Real-Time Electroanalytical Detection of Sulfide Via Laser-Induced Graphene Sensor


Zitieren

Introduction

Hydrogen sulfide (H2S) is not only considered harmful to health due to its toxicity but also has a bad smell like that of rotten egg. It plays a significant role in biological nerve regulation, vascular tone regulation, and reduction of metabolic state in the physiological system. Along with nitric oxide (NO) and carbon monoxide (CO), the endogenous H2S is considered to be the third-most crucial gasotransmitter [1, 2, 3, 4]. Several environmental regularities recommend monitoring the sulfide level as having an upper limit, and appropriate discharge levels are often imposed in the workplace. The toxicity of the liberated H2S is a problem for those handling sulfide-contaminated processes, such as in the petroleum industries, sewage treatments, and industrial discharge treatment [5, 6]. The unpleasant odor of H2S can be attributed to its neurotoxic effect and often contributes to severe, sometimes fatal, conditions due to sulfide poisoning [7, 8, 9]. Thus, it is significant to have a susceptible and selective, but user-friendly sensor to detect the hazardous gas in both gaseous and dissolved liquid phases [10].

The conventional sulfide-detection approach included the colorimetric method via the methylene blue (MB) test, initially reported by Enail Fisher, and having high selectivity and sensitivity. Other methods for the detection were the fluorescence method and the electrochemical technique using an amperometry-based device. [7, 8, 9]. The chemistry between MB and sulfide has been substantially proven, whereby the reaction of MB with sulfide involves oxidation, which led to the development of an electrochemical technique [12, 13]. Therefore, using MB in the electrochemical sensing of sulfide was an efficient approach. Electrochemical performances were significantly affected by the surface structure, composite, particle size, and shape. Thus, using a suitable nanoparticle, which increases the surface area and the active sites, was recommended to enhance the electrochemical signal [14, 15, 16, 17, 18]. However, external factors, such as temperature, concentration, and interfering substances may influence the stability of the sensor [19, 20]. Hence, immobilizing MB over a suitable nanomaterial matrix will enhance the electrocatalytic activity. The literature reports suggest using MB for sulfide sensing and other bulk electrodes as found in Refs. [12, 21]. Table 1 summarizes a few relevant reports, comparing them with the present work. Notably, most of the electrodes utilized in the reported works were bulky and require multistep fabrications.

Comparison with the previously reported works.

Material Method Range (μM) LoD (μM) Ref.
CoPCNF/GCE CV 75–770 46 [22]
CEC CV 100–1000 9 [23]
Mercury/Platinum CSSV 1–20 0.25 [24]
HMDE SWP 0.2–83 0.1 [25]
HMDE DPCSV 3–20 2 [26]
BDD CV 20–100 0.8 [27]
GCE SWV 3–120 0.10 [28]
LIG/MB CA/CV 0.5–500 0.435 This work

BDD, boron-doped diamond electrode; CA, chronoamperometry; CEC, carbon epoxy composite; CoPCNF, Cobalt pentacyanonitrosylferrate; CSSV, cathodic stripping sweep voltammetry; CV, cyclic voltammetry; DPCSV, differential pulse cathodic stripping voltammetry; GCE, glassy carbon electrode; HMDE, hanging mercury drop electrode; LIG, laser-induced graphene; MB, methylene blue; SWP, square wave polarography.

The present work reports a miniaturized, three-electrode system with laser-induced graphene (LIG) electrodes for the electrochemical sensing of sulfide. Herein, the working electrode was modified with MB [12]. This phenothiazine dye bonds firmly with LIG by a π–π bond, as LIG has a large surface area and good adsorption capability. Further, Ag/AgCl coated and bare LIG were used as reference and counter electrodes, respectively. Using cyclic voltammetry (CV) and chronoamperometry (CA) techniques, the fabricated sensor showed excellent interference-mitigated sensing of sulfide in a linear range of 0.5–500 μM with the LoD and LoQ being 0.435 μM, 2.45 μM, respectively. The sensor was also tested with other interfering analytes, such as sodium nitrite (NaNO2), sodium sulfide (Na2SO3), sodium carbonate (Na2CO3), dopamine (DA), and ascorbic acid (AA), and with the real lake water sample. This miniaturized sensor also showed good reproducibility and repeatability, leading to its potential for field-testing deployment. The challenge here was to obtain the reproducibility, as the analyte was prone to escape from the electrolyte. However, to overcome this, we have used the salt phase of sulfide. Further, the sensor miniaturization requires skilled coating, which makes it difficult to reproduce. Hence, a minimal, fixed volume of redox mediator (MB) was used to obtain the reproducibility.

Material and Method

Methylene blue (ISOCHEM, Kochi, India) was obtained from Amazon, sodium sulfide flakes (ISOCHEM, Kochi, India), sodium phosphate dibasic dehydrated (Na2HPO4.2H2O), and sodium phosphate monobasic dihydrated (NaH2PO4.2H2O) were obtained from sisco research laboratory (SRL), India. Sodium nitrite (NaNO2), sodium sulfite (Na2SO3), sodium arbonate (Na2CO3), DA, and AA were taken from Sigma Aldrich, India. For electrochemical analysis, Biologic Potentiostat (SP-150, France) was used. The polyimide sheet has been used as a substrate. CO2 laser (VLS 3.60 from Universal Laser System, Phoenix, AZ, USA) was used for the laser ablation process.

Modifications over the working electrodes

To modify the working electrode, 10 mg of MB was mixed in 10 ml of DI water using the ultrasonication bath for 1 h. Afterward, 1 μL of the MB solution was drop cast on the specified surface of the working electrode and kept in a desiccator overnight for the immobilizing of MB over LIG.

Method of sensor fabrication

The sensor was designed in computer aided design (CAD) software and converted to the .dxf file format, which was in turn converted to the .cdr file format. Further, it was processed using Coral draw compatible with the software of the CO2 laser system, which was used for the laser ablation process. For fabrication, a polyimide sheet was used, and the laser ablation process was done with a power of 6.5% of 30 W and a laser speed of 4.5% of 1200 mm/s. The sensor was fabricated after the laser ablation, as in Scheme I.

Scheme I:

Schematic of the electrochemical sensor fabrication with the real image of the sensor.

Result and Discussion
Physico-chemical analysis

The developed sensing material was thoroughly characterized to understand its characteristics better. Various characterizations were performed, such as scanning electron microscopy (SEM), X-ray diffraction (XRD), and X-ray photoelectron scattering (XPS).

SEM

SEM was leveraged to understand the surface morphology and the material structure's pore-size distribution. This implies that surface carbonization has been performed, which is required for the sensing. In Figure 1, the SEM images of different resolutions are provided. In addition, energy-dispersive X-ray analysis was performed to understand the material composition on the surface. Herein, different compositions of the material with their respective percentages are provided. The results indicate that the carbon has the very high conductivity and here in this work the laser induced graphene has been used which also has good conductivity as well as high surface area to volume ratio which is the required for the sensing. Furthermore, the surface modification carried out with MB can be seen. MB plausibly reacts with H2S for electrooxidation to take place [11].

Figure 1:

(a) SEM image of bare LIG (b) modified with MB (c) EDX analysis. EDX, energy dispersive X-ray; LIG, laser-induced graphene; MB, methylene blue; SEM, scanning electron microscopy.

XRD

The XRD of the sensing material has been carried out to understand the lattice effect of modifying the sensing material. Figure 2(a) is the XRD peak of the LIG/MB. The peak at 32° can be referred to as MB, the peak around 26° can be associated with the LIG, and the peak around 22° can be referred to as the polyimide sheet peak.

Figure 2:

(a) XRD of the LIG/MB (b) XPS of the LIG/MB modification. LIG, laser-induced graphene; MB, methylene blue; XRD, X-ray diffraction; XPS, X-ray photoelectron scattering.

XPS

For understanding bond formation, XPS is useful. Figure 2(b) is the XPS analysis of the LIG/MB working electrode. The respective element has a peak at the relevant binding energy, such as for carbon at 285.27 eV, nitrogen at 400.31 eV, oxygen at 533.05 eV, chlorine at 198.76 eV, and sulfur at 165.3 eV. The atomic weight percentages are as follows: carbon 81.1, nitrogen 3, oxygen 5.76, chlorine 0.18, and sulfur 0.76.

Electrochemical Characterization

To understand the electrochemical behavior of the working electrode, electrochemical characterization was carried out. Electrochemical impedance spectroscopy (EIS) was performed to monitor the surface modification's step-by-step effect on LIG. The impedance analysis includes a semicircle portion known as Warburg impedance [21]. Ferricyanide was used for the electrochemical characterization of the sensor. The initial and final frequency values were 10 mHz and 100 kHz, respectively, per the standard frequency adapted for the ferricyanide test [29, 30, 31, 32].

The electron transfer phenomena of [Fe(CN)6]3−/4− at different levels of surface modification of the electrode are shown in Figure 3(a). It was observed that the bare LIG has a larger Warburg area when compared to the MB-modified working electrode, which reduces Rct and enhances the charge transfer rate and sensitivity [21, 33]. Additionally, the CV response of the modified and unmodified electrodes was performed in a solution of 5 mM of K4[Fe(CN)6] and 1 mM of KCl dissolved in the Deionized water (DI) water type II at 50 mV/S for n = 4 cycle in the potential window of −1 V to +1 V.

Figure 3:

(a) EIS of bare LIG and LIG/MB, (b) Comparative CV response of LIG and LIG/MB with the potassium ferricyanide–KCl solution for 50 mV/S, n = 4 (number of cycle). CV, cyclic voltammetry; EIS, electrochemical impedance spectroscopy; LIG, laser-induced graphene; MB, methylene blue.

It is clear from Figure 3 that surface modification enhances the sensor performance. Figure 3(a) shows that the modification reduced the charge transfer resistance. At the same time, Figure 3(b) depicts that the modification enhances the current due to the increase in the surface-to-volume ratio.

Electrochemical Sensing of H2S using the LIG/MB–based electrode

The electrochemical experiment was performed using the MB-modified working electrode via CV, wherein the potential window was selected between −1 V to +1 V and the scan rate was 10 mV/s. Figure 4(a) shows a comparative CV of bare LIG in 0.1 M phosphate buffer solution (PBS) and 1 mM sodium sulfide (Na2S) solution. While the characteristic peak was observed, the MB-modified LIG shows a discrete oxidation peak at a potential of 0.17 V, as shown in Figure 4(b) with 1 mM Na2S at 10 mV/s for n = 4. MB was often used as the redox mediator for H2S detection [9].

Figure 4:

(a) Comparative CV of bare LIG in pH7 PBS and 1 mM Na2S at 10 mV/S for n = 4, (b) Comparative CV of LIG/MB in (pH = 7) PBS and 1 mM Na2S at 10 mV/S for n = 4. CV, cyclic voltammetry; LIG, laser-induced graphene; MB, methylene blue; PBS, phosphate buffer solution.

Scan rate effect

The scan rate analysis verified that the modification done using MB was successful, as the peak pertaining to MB was observed in the PBS due to electrochemical uptake. For this analysis, a buffer solution of 0.1 M PBS (pH = 7.4) was taken, and the scan rate was changed between 10 mV/s and 100 mV/s. As shown in Figure 5(a), a clear peak of MB has been observed. Further, the effect of varying scan rates over the electrocatalytic oxidation of LIG/MB over 1 mM Na2S was tested, as depicted in Figure 5(b) [34, 35].

Figure 5:

(a) Scan rate of 0.1 M PBS (pH = 7.4) on LIG/MB electrochemical sensor, (b) Scan rate test of 1 mM Na2S on LIG/MB electrochemical sensor, (c) Scan rate calibration curve (Ipa vs. sqrt(ν)) (d) Laviron plot of the LIG/MB electrochemical sensor from the scan rate plot. LIG, laser-induced graphene; MB, methylene blue; PBS, phosphate buffer solution.

As revealed in Figure 5(b), the redox peak was seen to increase gradually while enhancing the current. It can be seen that by using the Randle–Sevick equation, the current displays a linear relationship with the square root of the scan rate [19, 27, 28]. The linear plot shown in Figure 5(c) regarding the diffusion coefficient was 9 × 10−6 cm2 ⋅ s−1 [29]. To calculate the number of electrons that take part in the reaction, the Laviron plot was used, and the Laviron plot's slope gives the value. Ip=2.69×105×n×(α×nα)0.5×A×D0.5ν0.5CA {I_{\rm{p}}} = 2.69 \times {10^5} \times n \times {(\alpha \times {n_\alpha})^{0.5}} \times A \times {D^{0.5\nu 0.5}}{C_{\rm{A}}} where Ip indicates the peak current, n the number of electrons, α the charge transfer coefficient, nα the number of electrons involved in the reaction, D the diffusion coefficient of the analyte, A the area of the electrochemical sensor, and C the concentration of the analyte in bulk. The calculated electrochemical active surface area of the fabricated electrochemical sensor was 0.1495 mm2. The empirical Wilke–Chang formula can help calculate the diffusion coefficient [36].

D=7.4×108(χM)0.5T/(ν×V0.6) D = 7.4 \times {10^{- 8}}\,{(\chi M)^{0.5}}\,T/(\nu \times {V^{0.6}})

Where T indicates the temperature, M the molecular weight of the solvent, χ the association parameter (2.6 in the present case), ν the viscosity at 25°C, and V the molar volume of the solute at the boiling point.

The Laviron equation: Epa=E°+RTαnFlnRTKsαnF+RTαnFlnv {E_{{\rm{pa}}}} = E^\circ + {{RT} \over {\alpha nF}}ln{{RTKs} \over {\alpha nF}} + {{RT} \over {\alpha nF}}lnv where α represents the transfer coefficient, Ks the standard rate constant, n the electron's number, E° the redox potential, T the absolute temperature, and F the Faraday constant. Thus, Epa is proportional to lnν. The slope value was 1.32, αn = slope of the Epa vs. lnν = 1.32, and α = 0.5. This leads to the value of n as 2.64, concluding that the number of electrons in the reaction was 3 [33].

Concentration effect

Further, to understand the device performance, it was necessary to perform the concentration effect analysis in which the sensor detection limit was determined. The device performance was analyzed using the CA technique in 0.1 M PBS (pH = 7.4). Each concentration was tested for 4 min in the potential value of 0.17 V obtained by the CV. While performing the CA, there was an oxidation of S2− to S0 on the surface of the LIG/MB electrochemical sensor.

The concentration effect was performed thrice (n = 3) to ensure the accuracy and repeatability of the developed sensor. Then, the mean value and error bar were taken as the response values for the calculation. Figure 6 shows the concentration effect analysis in the 0.5 μM–1 mM Na2S range. It was observed that as the concentration was decreasing, the current response was also decreasing. This was due to the buildup of sulfur on the surface of the working electrode and the leaching of MB from the surface of the working electrode.

Figure 6:

(a) Concentration analysis on LIG/MB–based electrochemical sensor (b) calibration curve of the concentration analysis Ipa vs. concentration. Two linear ranges can be observed. LIG, laser-induced graphene; MB, methylene blue.

There were two linear ranges observed from the calibration curve of the current vs. concentration, 0.5–50 μM and 100 μM–1 mM, and their respective linear equations are as follow: Ip=0.0041×concentration+0.181,R2=0.969(0.550μM)Ip=0.0007×concentration+0.4964,R2=0.957(100μM1mM) \matrix{{{I_{\rm{p}}} = 0.0041 \times {\rm{concentration}} + 0.181,} \hfill \cr {{R^2} = 0.969(0.5 - 50\mu {\rm{M}})} \hfill \cr {{I_{\rm{p}}} = 0.0007 \times {\rm{concentration}} + 0.4964,} \hfill \cr {{R^2} = 0.957\,(100\,\mu {\rm{M}} - 1\,{\rm{mM}})} \hfill \cr}

The LoD was calculated to be 0.435 μM, and the sensitivity of the two linear ranges (low and high) for the electrochemical sensor was calculated as Sensitivity=0.295μA/(μMmm2)for0.550μMSensitivity=0.0047μA/(μMmm2)for1001mM \matrix{{{\rm{Sensitivity}} = 0.295\mu {\rm{A}}/(\mu {\rm{M}}\,{\rm{m}}{{\rm{m}}^2})\,{\rm{for}}\,0.5 - 50\,\mu {\rm{M}}} \hfill \cr {{\rm{Sensitivity}} = 0.0047\mu {\rm{A}}/(\mu {\rm{M}}\,{\rm{m}}{{\rm{m}}^2})\,{\rm{for}}\,100 - 1\,{\rm{mM}}} \hfill \cr}

The developed sensor's signal-to-noise (S/N) ratio was found to be 2.76. The limit of quantification (LoQ) was found to be 2.45 μM.

pH effect

The pH dependency of the solvent was a significant parameter for the oxidation reaction, as it can affect the reaction between LIG/MB and sulfide. Here the pH range was taken between 3 and 11 to study the behavior of the electrocatalytic oxidation in various acidic and basic pH ranges. The value of the Epa was plotted against the respective pH value. It shows that the electrochemical sensor was pH dependent. The number of protons involved in the reaction can be measured using the Nernst equation;

Epa=E0RTnFln[O][R]2.303mRTnFpH {E_{{\rm{pa}}}} = {E_0} - {{RT} \over {nF}}ln{{[O]} \over {[R]}} - 2.303{{mRT} \over {nF}}{\rm{pH}}

Through the pH analysis, it can be observed in Figure 7 that as the pH of the buffer solution varied from low to high, there was a shift in the peak potential and peak current. The respective Epa vs. pH and Ipa vs. pH, is shown in Figure 7(b, c). Thus, it can be observed that for each pH, a different number of protons was involved in the reaction and thus the pH of the PBS plays and important role in the sensing. As the charge transfer rate changes, the detection also gets affected [25].

Figure 7

(a) pH analysis of PBS on the LIG/MB electrochemical sensor ranging from 3 to 11 pH (b) Ipa vs. pH (c) Epa vs. pH. LIG, laser-induced graphene; MB, methylene blue; PBS, phosphate buffer solution.

Interference effect

Interference tests have been performed to determine the electrochemical sensor's selectivity. For this, various interfering analytes such as NaNO2, Na2SO3, Na2CO3, DA, and AA with 100 μM concentration were taken and mixed in 5 μM concentration of Na2S predissolved in 0.1 M PBS (pH = 7.4). In addition, O2 and CO2 were purged in the Na2S solution for 10 min to vary the effect of gases. It can be seen that there was no significant interference of the various analytes mentioned above, and the sensor shows good selectivity. The current change mentioned in percentage at the top of the bar graph and the result of the interference study are displayed in Figure 8

Figure 8:

Interference effect analysis (a) with the direct interfering analyte (b) with the gases.

Repeatability, reproducibility, and stability

The repeatability of the device has been analyzed and the respective bar graph is shown in Figure 9(a). The relative standard deviation (RSD) was found to be 2.62% for the repeatability of the sensor as can be seen in Figure 9(a). For the reproducibility of the sensor, three different sensors have been fabricated and tested for sulfide detection. The bar graph has been shown in Figure 9(b) along with the error bar and the respective RSD is 2.26%. This analysis indicates that the fabricated sensor shows good repeatability and reproducibility for sensing applications. For stability analysis, the device was used for 15 days The fabricated electrochemical sensor shows good stability as the deviations were around 3.67% for the peak potential value and 4.24% for the peak current value, respectively.

Figure 9:

(a) Repeatability and (b) reproducibility of the fabricated electrochemical.

Real sample analysis

Real sample testing was performed to verify the fabricated electrochemical sensor's practical ability. The lake water sample was collected from different lakes, such as Kapra Lake, Hussain Sagar Lake, and Shameerpet Lake in Hyderabad, India. For the real sample analysis, the sample preparation involved mixing the sample with 10-fold of 0.1 M PBS (pH = 7.4) solution. Further, different concentrations of the Na2S, such as 10 μM, 50 μM, and 100 μM, were used in the sample prepared. The electrochemical technique used was CA at a potential of 0.17 V and was run for 5 min. Following this, recovery analysis was performed in which the electrochemical sensor showed good recovery. The low RSD indicates that the device performance is good. The result of the real sample analysis is given in Table 2. Also, the respective CA was provided for each lake with different concentrations spiked in Figure 10(a–c).

Real sample analysis.

Source of Lake Water S. No. Added (μM) Found (μM) Recovery (%) RSD
Shameerpet Lake 1 10 9.866667 98.7 2.055076
2 50 49.26667 98.6 4.72582
3 100 99.3 99.3 2.64575
Hussain Sagar Lake 1 10 9.8 98 1.37477
2 50 49.71333 99.44 2.41316
3 100 99.89 99.89 3.6056
Kapra Lake 1 10 9.843333 98.5 1.76376
2 50 49.07667 98.2 4.45459
3 100 99.61 99.61 2.91376

RSD, relative standard deviation.

Figure 10:

Real Sample analysis of (a) Kapra Lake, (b) Hussain Sagar Lake, (c) Shameerpet Lake.

Conclusion

The present study reports the fabrication of a miniaturized electrochemical sensor for detecting sulfide. A miniaturized three-electrode system with LIG electrodes has been fabricated, where one LIG electrode was modified with MB and used as a working electrode. The LIG electrode modified with Ag/AgCl ink was used as the reference electrode, whereas bare LIG electrodes were used as counter electrodes. The electrode demonstrates excellent electroactive oxidation of sulfide at E0 = 0.17 V. The modified system has good linearity and sensitivity with a linear range of 0.5–500 μM, LoD of 0.435 μM, and LoQ of 2.45 μM. The pH effect was performed to select the optimum pH of the buffer solution. The fabricated electrode shows promising results in the monitoring of lake water. The sensor demonstrated good sensitivity, selectivity, and stability for 15 days. In future, the integration with the microfluidic based device for the flow and droplet-based study can be carried out.

There are no sources in the current document.

eISSN:
1178-5608
Sprache:
Englisch
Zeitrahmen der Veröffentlichung:
Volume Open
Fachgebiete der Zeitschrift:
Technik, Einführungen und Gesamtdarstellungen, andere